B 97505
B 97505
Volume 23
Editors
S.S. Antman J.E. Marsden
L. Sirovich S. Wiggins
The purpose of this series is to meet the current and future needs for the interac-
tion between various science and technology areas on the one hand and mathe-
matics on the other. This is done, firstly, by encouraging the ways that mathe-
matics may be applied in traditional areas, and well as point towards new and
innovative areas of applications; and, secondly, by encouraging other scientific
disciplines to engage in a dialog with mathematicians outlining their problems to
both access new methods and suggest innovative developments within mathe-
matics itself.
The series will consist of monographs and high-level texts from researchers
working on the interplay between mathematics and other fields of science and
technology.
Interdisciplinary Applied Mathematics
Volumes published are listed at the end of this book.
Springer
New York
Berlin
Heidelberg
Hong Kong
London
Milan
Paris
Tokyo
Muhammad Sahimi
Heterogeneous Materials
Nonlinear and Breakdown Properties
and Atomistic Modeling
Disorder plays a fundamental role in many natural and man-made systems that
are of industrial and scientific importance. Of all the disordered systems, hetero-
geneous materials are perhaps the most heavily utilized in all aspects of our daily
lives, and hence have been studied for a long time. With the advent of new ex-
perimental techniques, it is now possible to study the morphology of disordered
materials and gain a much deeper understanding of their properties. Novel tech-
niques have also allowed us to design materials of morphologies with the properties
that are suitable for intended applications.
With the development of a class of powerful theoretical methods, we now have
the ability for interpreting the experimental data and predicting many properties
of disordered materials at many length scales. Included in this class are renor-
malization group theory, various versions of effective-medium approximation,
percolation theory, variational principles that lead to rigorous bounds to the ef-
fective properties, and Green function formulations and perturbation expansions.
The theoretical developments have been accompanied by a tremendous increase in
the computational power and the emergence of massively parallel computational
strategies. Hence, we are now able to model many materials at molecular scales
and predict many of their properties based on first-principle computations.
In this two-volume book we describe and discuss various theoretical and com-
putational approaches for understanding and predicting the effective macroscopic
properties of heterogeneous materials. Most of the book is devoted to comparing
and contrasting the two main classes of, and approaches to, disordered materials,
namely, the continuum models and the discrete models. Predicting the effective
properties of composite materials based on the continuum models, which are based
on solving the classical continuum equations of transport, has a long history and
goes back to at least the middle of the nineteenth century. Even a glance at the liter-
ature on the subject of heterogeneous materials will reveal the tremendous amount
of work that has been carried out in the area of continuum modeling. Rarely, how-
ever, can such continuum models provide accurate predictions of the effective
macroscopic properties of strongly disordered multiphase materials. In particular,
if the contrast between the properties of a material’s phases is large, and the phases
form large clusters, most continuum models break down. At the same time, due to
their very nature, the discrete models, which are based on a lattice representation
of a material’s morphology, have the ability for providing accurate predictions for
the effective properties of heterogeneous materials, even when the heterogeneities
are strong, while another class of discrete models, that represent a material as a
collection of its constituent atoms and molecules, provides accurate predictions of
viii Preface
the material’s properties at mesoscopic scales, and thus, in this sense, the discrete
models are complementary to the continuum models. The last three decades of
the twentieth century witnessed great advances in discrete modeling of materials
and predicting their macroscopic properties, and one main goal of this book is to
describe these advances and compare their predictions with those of the continuum
models. In Volume I we consider characterization and modeling of the morphology
of disordered materials, and describe theoretical and computational approaches for
predicting their linear transport and optical properties, while Volume II focuses
on nonlinear properties, and fracture and breakdown of disordered materials, in
addition to describing their atomistic modeling. Some of the theoretical and com-
putational approaches are rather old, while others are very new, and therefore we
attempt to take the reader through a journey to see the history of the development
of the subjects that are discussed in this book. Most importantly, we always com-
pare the predictions with the relevant experimental data in order to gain a better
understanding of the strengths and/or shortcomings of the two classes of models.
A large number of people have helped me gain deeper understanding of the
topics discussed in this book, and hence have helped me to write about them.
Not being able to name them all, I limit myself to a few of them who, directly
or indirectly, influenced the style and contents of this book. Dietrich Stauffer has
greatly contributed to my understanding of percolation theory, disordered media,
and critical phenomena, some of the main themes of this book; I am deeply grateful
to him. For their tireless help in the preparation of various portions of this book, I
would like to thank two of my graduate students, Sushma Dhulipala and Alberto
Schroth. Although they may not be aware of it, Professors Pedro Ponte Castañeda
of the University of Pennsylvania and Salvatore Torquato of Princeton University
provided great help by guiding me through their excellent work, which is described
in this book; I would like to thank them both. Some of my own work described in
this book has been carried out in collaboration with many people; I am pleased to
acknowledge their great contributions, especially those of Dr. Sepehr Arbabi, my
former doctoral student. The constant encouragement and support offered by many
of my colleagues, a list of whom is too long to be given here, are also gratefully
acknowledged. I would like particularly to express my deep gratitude to my former
doctoral student Dr. Jaleh Ghassemzadeh, who provided me with critical help at
all stages of preparation of this book. Several chapters of this book have been used,
in their preliminary versions, in some of the courses that I teach, and I would like
to acknowledge the comments that I received from my students.
My wife, Mahnoush, and son, Ali, put up with the countless hours, days, weeks,
and months that I spent in preparing this book and my almost complete absence
during the time that I was writing, but never denied me their love and support
without which this book would have never been completed; I love and cherish
them both.
Muhammad Sahimi
Los Angeles, California, USA
May 2002
Contents
Preface vii
Introduction to Volume II 1
References 593
Index 633
Abbreviated Contents for Volume I
Preface
1 Introduction
I Characterization and Modeling of the Morphology
2 Characterization of Connectivity and Clustering
3 Characterization and Modeling of the Morphology
II Linear Transport and Optical Properties
4 Effective Conductivity, Dielectric Constant, and
Optical Properties: The Continuum Approach
5 Effective Conductivity and Dielectric Constant:
The Discrete Approach
6 Frequency-Dependent Properties: The Discrete Approach
7 Rigidity and Elastic Properties: The Continuum Approach
8 Rigidity and Elastic Properties: The Discrete Approach
9 Rigidity and Elastic Properties of Network Glasses, Polymers,
and Composite Solids: The Discrete Approach
References
Index
Introduction to Volume II
A. Constitutive Nonlinearity
Materials of this type always behave nonlinearly. For example, if in a composite
material the relation between the current I and voltage V is given by
I = gV n
B. Threshold Nonlinearity
In this class of materials are those for which the nonlinearity arises as a result
of imposing on them an external field of sufficient intensity. Brittle fracture and
dielectric breakdown of composite solids are two important examples of such
nonlinear transport processes. In brittle fracture, for example, the elastic response
of a solid material is governed by the equations of linear elasticity until the external
stress or strain that has been imposed on the material exceeds a critical value, at
which time the material breaks down and microcracks begin to emerge. A list
of all possible nonlinear transport processes of this type is very long. This type
of nonlinearity will be studied in Chapters 5–8, and will include electrical and
dielectrical breakdown, brittle fracture, and the transition between brittle fracture
and ductile behavior.
One important point to remember is that, the interplay between a nonlinear
transport process and the disordered morphology of a composite material gives
rise to a rich variety of phenomena that are usually far more complex than what one
usually must deal with in linear processes. Over the past 15 years, an increasing
number of investigations have been devoted to such nonlinear transport processes,
and deeper insight into their properties has been acquired. A major goal of Volume
II is to describe this progress and compare various properties of nonlinear transport
processes in heterogeneous materials with their linear counterparts.
C. Theoretical Approaches
Although the analysis of transport processes in composite materials has a long
history, it is only in the past three decades that this analysis has been extended
to include detailed structural properties of the materials, and in particular the
distribution of their heterogeneities. Deriving exact results for the effective prop-
erties of composite materials with anything but the simplest morphologies is
extremely difficult, if not impossible, and thus one must resort to various approx-
imate techniques. At the same time, however, the advent of powerful computers
and development of efficient computational algorithms have allowed us to esti-
mate various properties of heterogeneous materials to practically any desired or
affordable accuracy.
To describe the theoretical approaches for estimating the effective properties of
composite materials, we divide them into two classes. In the first class of models are
what we refer to as the continuum models, while the second class is made of the dis-
crete models. Both types of models are described and analyzed in this Volume, and
what follows is a brief description of the general features of each class of models.
sible computations for exact estimation of the effective properties are still very
difficult—even in the event that one knows the detailed morphology of the material.
Thus, it becomes essential to adopt a macroscopic description at a length scale much
larger than the dimension of the individual phases of a composite material. The
governing equations are then discretized and solved numerically, provided that the
effective properties that appear in the transport equations are either supplied as the
inputs (through, for example, experimental measurements), or else a model for the
morphology of the material is assumed so that the effective transport properties
can be somehow estimated, so that the numerical solution yields other quantities
of interest, such as the potential distribution in the material. We refer to various
models associated with this classical description as the continuum models. These
models have been widely used because of their convenience and familiarity to the
engineers and materials scientists. Their limitations will be described and discussed
in the subsequent chapters.
In addition to deriving the effective macroscopic equations and obtaining their
solution by numerical calculations, one may also derive exact results in terms of
rigorous upper and lower bounds to the properties of interest. Hence, powerful
tools have been developed for deriving accurate upper and lower bounds and
estimates. Finally, various approximations, such as the mean-field and effective-
medium approximations, have also been developed in the context of the continuum
models. We will describe most of these theoretical approaches throughout both this
book and Volume II.
The main shortcoming of both groups of the discrete models, from a practical
point of view, is the large computational effort required for a realistic discrete repre-
sentation of the material and simulating its behavior, although the ever-increasing
computational power is addressing this difficulty.
1.0 Introduction
Natural, as well as man-made, materials have enormous variations in their mor-
phology, which consists of materials’ geometry, topology and surface structure.
The geometry refers to sizes of the micro- and mesoscale elements of the mate-
rials, as well as their shapes which range anywhere from completely ordered to
complex and seemingly chaotic patterns. Generally speaking, regular Euclidean
shapes are formed under close-to-equilibrium conditions, although even in such
cases equilibrium thermodynamics is often incapable of describing the process
that gives rise to such shapes. The topology of materials describes how the micro-
and mesoscale elements are connected to one another. The structure of materials’
surface, especially those that are produced under far-from-equilibrium conditions,
is also very important because the surface is often very rough and possesses com-
plex features. In recent years it has become clear that characterizing the surface
roughness will go a long way toward giving us a much better understanding of
materials’ microstructure and hence many of their effective properties. However,
when we speak of surface roughness, we must specify the length scales over
which the roughness is measured. Even the most rugged mountains look perfectly
smooth when viewed from the outer space! Therefore, surface roughness (and,
more generally, all the morphological characteristics) depends on the length scale
of observations or measurements. The effect of topology of disordered materials
on their effective transport properties is quantified by percolation theory which,
together with the effect of the geometry, was described in Chapters 2 and 3 of
Volume I, and their significance was emphasized throughout Volume I where we
analyzed effective linear properties of disordered materials. For other applications
of percolation theory see Sahimi (1994a). Stauffer and Aharony (1992) present a
simple introduction to the concepts of percolation theory. In this chapter, we con-
sider the structure and characteristics of materials’ surface, and describe various
theoretical and experimental methods of studying rough surfaces, which are di-
rectly relevant to the nonlinear phenomena in heterogeneous materials considered
in this Volume, particularly to their brittle fracture and dielectric breakdown.
An important example of a material with a rough surface are the thin films
that produced by molecular beam epitaxy, and are utilized for manufacturing of
semiconductors and computer chips. These films are made of silicon and other
6 1. Characterization of Surface Morphology
elements, and are prepared by deposition of atoms on a very clean surface. Thin
films with rough surfaces are also made by sputtering in which an energized beam
of particles is sent toward the bulk of a material. Collision of the beam particles
with the material causes ejection of some particles from the material’s surface,
which then deposit on another surface and start to grow a thin film of the original
material.
Although the enormous variations in the morphology of natural, and even man-
made, materials, particularly in their surface, are such that, up until a few decades
ago, the problem of describing and quantifying such morphologies seemed hope-
less, many experimental and theoretical developments of the past two decades have
brightened the prospects for deeper understanding of materials’ microstructures,
and in particular the structure of their surface. Among them are the advent of pow-
erful computers and novel experimental techniques that allow highly sophisticated
computations of materials’ properties and their measurement. In addition, the real-
ization that the complex microstructure and behavior of a wide variety of materials
can be quantitatively characterized by utilizing the ideas of fractal distributions,
have advanced our understanding of materials’ surface structure. As we discuss in
this chapter, fractal concepts provide us with a powerful tool for characterizing the
structure of materials’ surface and its roughness, and the long-range correlations
that often exist in their morphology.
The purpose of this chapter is to describe and discuss the essential features of
surface morphology and its dynamics during the process in which it is formed, and
how fractal concepts can be utilized for characterizing it. We already described in
Chapter 2 of Volume I most of the main concepts of fractal geometry, and therefore
in this chapter we restrict ourselves to a brief discussion of such concepts, after
which we study and analyze rough surfaces.
where s(r) is a function such that s(r) = 1 if a point at r belongs to the system,
s(r) = 0 otherwise, and r = |r|. Because of self-similarity of the system, the direct
1.3. Rough Surfaces: Self-affine Fractals 9
but has very little effect in the other directions, hence generating anisotropy in the
structure of rock. The interested reader is referred to Family and Vicsek (1991) for
an excellent collection of articles which describe a wide variety of rough surfaces
with self-affine properties.
Self-affine fractals that one encounters in practical situations are typically disor-
dered, and thus their self-affinity is only in a statistical sense. For the problems that
are of interest to us in this book, a disordered self-affine fractal can be thought of as
the fluctuations about a straight line or a flat surface. Such fluctuations can generate
rough self-affine curves or surfaces. If we consider the height difference between a
pair of points h(x1 ) and h(x2 ) on a self-affine surface h(x) that lie above or below
points separated by a distance x1 − x2 = x = |x| on a flat reference surface (or
line), then
|h(x1 ) − h(x2 )| ∼ x H , (9)
where H is called the Hurst exponent. One may generalize Eq. (9) to higher dimen-
sions, and generate rough surfaces that are encountered in a variety of contexts,
from surface of pores of a natural porous medium (see, for example, Sahimi, 1993b,
1995b, for comprehensive discussions) to fracture surface of heterogeneous ma-
terials (see Chapters 6 and 7), to thin films that are formed by a deposition process
(see below).
Figure 1.2 presents 1D and 2D rough profiles and surfaces generated by fBm.
The increments in fBm are stationary but not ergodic. The variance of a fBm
for a large enough array is divergent (i.e., the variance increases with the size of
the array without bounds). Its trace in d dimensions is a self-affine fractal with
a local fractal dimension Df = d + 1 − H . Fractional Brownian motion is not
differentiable at any point, but by smoothing it over an interval one can obtain its
approximate numerical derivative which is called fractional Gaussian noise (fGn),
a 1D example of which is shown in Figure 1.3, which should be compared with its
counterpart in Figure 1.2. We should point out that the correlation function C(r)
of a fBm is given by
Figure 1.2. Examples of one- and two-dimensional rough profiles and surfaces generated
by the fractional Brownian motion with various Hurst exponents H .
The cutoff co allows us to control the length scale over which the spatial properties
of a system are correlated (or anticorrelated). Thus, for length scales L < co
the properties preserve their correlations (anticorrelations), but for L > co they
become random and uncorrelated. Note that the power spectrum of fGn in, for
1.4. Generation of Rough Surfaces: Fractional Brownian Motion 13
example, 1D is given by
bd
S(ω) = , (18)
ω2H −1
where bd is another d-dependent constant. The spectral representation of fBm (and
fGn) provides a convenient method of generating an array of numbers that follow
the fBm statistics, using a fast Fourier transformation (FFT) technique. In this
method, one first generates random numbers, distributed either uniformly in [0,1),
or according to a Gaussian distribution with random phases, and assigns them to
the sites of a d-dimensional lattice. In most cases the linear size L of the lattice is a
power of 2, but the only requirement is that L can be partitioned into small prime
numbers, so that a FFT algorithm can be used. One must also keep in mind that,
since the variance σ 2 of a fBm increases with the size L of the array, generating a
fBm array with a given variance requires selecting an appropriate L. In any case,
the Fourier transformation of the resulting d-dimensional array of the√numbers is
then calculated numerically, the resulting numbers are multiplied by S(ω), and
the results then inverse Fourier transformed back into the real space. The array so
obtained follows the statistics of a fBm with the desired long-range correlations
and the specified value of H . To avoid the problem associated with the periodicity
of the numbers arising as a result of their Fourier transforming, one must generate
14 1. Characterization of Surface Morphology
the array using a much larger lattice size than the actual size that is to be used in
the analysis, and use the central part of the array (or lattice).
ensure accuracy. For example, in our own work we have used up to 140 terms in
−70 ≤ j ≤ 70 to obtain accurate results. The power spectrum of the data array
generated by the WM method is discrete and does not contain all the frequencies.
However, it is still proportional to ω−(2H +1) , in agreement with Eq. (16).
If the rough self-affine fractal surface is growing with the process time t as in,
for example, deposition on a flat surface, then one must define a more general
correlation Cn (x, t) in a manner similar to that used for Cn (x), namely,
Cn (x, t) = [|h(x0 + x, t + t ) − h(x0 , t )|]n
1/n
, (26)
t .
where the averaging is over all the initial position x0 and times Then, due to
self-affinity of the surface, the correlation function Cn (x, t) has the property that
Cn (bx, bz t) = bα Cn (x, t). (27)
Similar to C2 (x), one usually constructs C2 (x, t) and attempts to extract from it
information about the surface. Under the dynamic conditions in which a rough and
self-affine surface grows, there exists a time scale tc over which the time correla-
tions are important. For rough surfaces that begin growing from a smooth surface,
it has been found in most cases that ξ⊥ and ξ satisfy the following power laws,
ξ⊥ ∼ t β , t tc (28)
ξ ∼t 1/z
, t tc (29)
where t is either the time (for a growing rough surface) or the surface’s mean thick-
ness. For t tc the magnitude of ξ saturates, ξ = L. The quantity z is called
the dynamical exponent of the surface, while β is called the growth exponent. The
quantities ξ⊥ and ξ are actually related to each other by
ξ⊥ ∼ ξ α . (30)
α is called the roughness exponent. Although we are not aware of an experimental
realization of a case for which α and the Hurst exponent H are different, we keep
both α and H to make our discussion as general as possible.
The roughness of a dynamic, growing surface is characterized by the width
w(L) defined as,
1/2
w(L) = [h(x) − hL ]2 , (31)
where h(x) is, as before, the height of the surface at position x, and hL is its
average over a horizontal segment of length L (normalized by the “volume” Ld−1 ).
According to the dynamic scaling theory of Family and Vicsek (1985) for growing
rough surfaces, one has the following dynamic scaling equation
h(x) − hL ∼ t β f (x/t β/α ), (32)
where α and β, the two exponents defined above, satisfy the following scaling
relation
α
α + = 2, (33)
β
and the scaling function f (u) has the properties that |f (u)| < c for u 1, and
f (u) ∼ Lα f (Lu) for u 1, where c is a constant. Note that the ratio α/β can be
replaced by the dynamical exponent z. It is then straightforward to show that
w(L, t) ∼ Lα g(t/Lα/β ), (34)
1.5. Scaling Properties of Rough Surfaces 17
where g(u) is a universal scaling function. Note also that w(L, t) is a measure of
the correlation length ξ⊥ along the direction of growth. As the rough surface grows,
the wavelength of the spatial fluctuations and the length over which the fluctuations
are correlated both grow with time. However, the length L is the maximum spatial
extent to which the correlations can grow in the d − 1 dimensions along the surface.
When the correlations reach this scale, they cannot extend further, and therefore
the rough surface reaches a steady-state which is characterized by a constant width.
Then, the surface is scale invariant and the saturation value w(L, ∞) is expected
to have a power-law dependence on L:
w(L, ∞) ∼ Lα . (35)
The correlation time tc also scales with L as
tc ∼ Lα/β ∼ Lz , (36)
Equation (34) indicates that, if one plots w/Lα versus t/Lα/β , then, due to the
universality of g(u), all the results for various t and L should collapse onto a single
universal curve [representing the scaling function g(u)]. Figure 1.4 presents such
a data collapse for a rough surface grown by a ballistic deposition process (Vold,
1963). In the simplest version of ballistic deposition, one begins with a line of
L, selects at random a horizontal line above the line of particles, and places a
α
Figure 1.4. Data collapse for a rough surface grown by ballistic deposition (courtesy of
Ehsan Nedaaee Oskoee).
18 1. Characterization of Surface Morphology
Figure 1.5. An example of a rough surface grown by ballistic deposition (courtesy of Ehsan
Nedaaee Oskoee).
particle there. The particle is then allowed to fall along a straight line vertically
downward. When the particle touches the original particles, it sticks to them and
becomes part of the particle pile. A large deposit is then grown by repeating this
procedure. Extensive numerical simulations indicate that the deposit is compact
and non-fractal, but its surface is rough and self-affine. An example is shown in
Figure 1.5.
the growth of the rough surface. However, because of various constraints that are
imposed by the physics of growing a rough surface, the set of acceptable functions
R(x, t) is limited. Some of these constraints are as follows (Barabási and Stanley,
1995).
(1) The growth of the surface should be independent of where h = 0 is defined,
i.e., it should be invariant under the transformation h → h + δh. Therefore,
R cannot depend explicitly on h, but should be built from such terms as ∇ n h
(with n = 1, 2, · · ·).
(2) The equation must have rotation and inversion symmetry with respect to the
direction of the growth, implying that it cannot contain odd-order derivatives
in the coordinates, such as ∇h and ∇(∇ 2 h).
(3) The equation must be invariant under time translation t → t + δt, which
means that R cannot depend explicitly on t. It should also be translationally
invariant in the direction perpendicular to the growth direction, and therefore
R cannot contain terms that are explicit in x.
(4) Since the fluctuations in the rough surface must be similar with respect to
the mean position of the surface—the so-called up-down symmetry (h → −h
invariance)—the equation cannot contain terms such as (∇h)n with n being an
even number. However, this symmetry can be broken if there exists a driving
force F, perpendicular to the rough surface, which selects a particular direction
for the growth of the surface. The existence of this driving force is a necessary
but not sufficient condition for breaking this symmetry.
Therefore, the most general form of the equation that describes the growth of a
rough surface is given by
∂h
= ∇ 2 h + ∇ 4 h + · · · + (∇ 2 h)(∇h)2 + · · · + (∇ 2k h)(∇h)2j + N (x, t).
∂t
(38)
To investigate the scaling properties of a growing surface, we consider the
hydrodynamic limit, t → ∞ and x → ∞. In this limit, the higher-order deriva-
tives of h are much smaller than the lowest-order one. Consider, as examples,
∇ 2 h and ∇ 4 h. Writing x → x ≡ bx, we must have h → h ≡ bα h, and thus,
∇ 2 h → ∇ 2 h ≡ bα−2 ∇ 2 h and ∇ 4 h → ∇ 4 h ≡ bα−4 ∇ 4 h. In the limit b → ∞,
∇ 4 h decays much faster than ∇ 2 h and can therefore be neglected.
Given such considerations, the simplest possible equation has the following
form
∂h
= D∇ 2 h + N (x, t), (39)
∂t
which was proposed by Edwards and Wilkinson (1982). In most cases, the noise
term has been assumed to be Gaussian:
N (x, t)N (x , t) = 2Aδ(x − x )δ(t − t ), (40)
where A is the amplitude of the noise. Equation (40) implies that there is no
correlation in space or time, since the average N (x, t)N (x , t) vanishes (except,
20 1. Characterization of Surface Morphology
Equation (48) suggests that the interface between the two materials, i.e., between
the rough surface and the coating material, in the cross-sections parallel to the
reference plane is a self-similar fractal with a fractal dimension
Df = d − H, (49)
where d is the Euclidean dimensionality of the reference surface. Therefore, if the
fractal dimension Df can be estimated independently, then the Hurst exponent
H can also be evaluated. Typically, the islands that appear have a surface area
distribution nS such that
nS ∼ S −τ , (50)
where nS is the number of islands with areas S in the range [S − 1
2 S, S+ 1
2 S].
The exponent τ is related to the fractal dimension Df through the following
equation
1
τ= Df + d , (51)
d
so that measurement of the islands’areas yields Df , from which the Hurst exponent
H can be estimated.
The third method of analyzing a rough, self-affine surface is through its power
spectrum which, in d dimensions, is given by Eq. (16). However, as Hough (1989)
pointed out, interpreting a power-law power spectrum is not without difficulties,
and thus one must be careful in using such an analysis. In particular, a power-law
power spectrum might also be the result of a non-stationary and non-fractal system.
We will come back to this issue in Chapters 6 and 7, where we describe fracture
surface of materials which are typically very rough.
Summary
An important characteristics of morphology of disordered multiphase materials
is the structure of their surface, and in particular their surface roughness. The
concepts of modern statistical physics of disordered media can now quantify the
roughness in terms of self-affine fractals, and the roughness or Hurst exponent.
The dynamics of growth of such surfaces can also be described by dynamical
scaling, discrete models of material growth, and suitable continuum differential
equations. Moreover, fractal geometry, and the associated power-law correlation
functions, point to the fundamental role of length scale and long-range correlations
in the macroscopic homogeneity of a heterogeneous material. If the largest relevant
length scale of the material, e.g., its linear size, is less than the length scale at
which it can be considered homogeneous, then the classical equations that describe
transport processes in the material must be fundamentally modified.
Part I
Effective Properties of
Heterogeneous Materials with
Constitutive Nonlinearities
2
Nonlinear Conductivity and Dielectric
Constant: The Continuum Approach
2.0 Introduction
The main focus of Volume II is on nonlinear properties of heterogeneous mate-
rials. In general, there are two fundamental classes of nonlinearity that one may
encounter in disordered materials:
(1) One class of nonlinear materials is described by what we call constitutive non-
linearity, which is one in which the basic local constitutive law that expresses
the relation between the flux (of current, force, etc.) and the potential (volt-
age, stress, etc.) gradient is nonlinear. As a result, the macroscopic behavior
of such materials must also be described by nonlinear transport equations. In
particular, the effective transport properties of such materials are nonlinear
in the sense of being functions of the external potential gradient. Such mate-
rials are of great practical importance, since, for example, one may be able
to design new nonlinear optical materials by tuning their nonlinear response
which can be achieved by, for example, changing the volume fraction of their
constituents. For example, it has been suggested that strong local field effects,
such as the large local field at the surface plasmon resonance frequency of a
metallic inclusion, may lead to enhanced nonlinear response in a heteroge-
neous material. Constitutive nonlinearity is the subject of this and the next
two chapters. Even within this restricted class of nonlinear materials, one may
imagine a very large number of nonlinear constitutive equations (similar to
those that have been proposed, for example, for polymeric fluids). Therefore,
while we describe in this chapter results for general constitutive nonlinearity,
their application is restricted mostly to strongly nonlinear materials, i.e., those
that are described by a power-law relation between the flux and the current.
In the next two chapters we will also describe the macroscopic behavior of
nonlinear materials that can be described by a few other types of nonlinear
constitutive equations, for which considerable progress has been made, and a
comparison between the theoretical predictions and the experimental data is
possible.
(2) In the second class of nonlinearities, a material is characterized by thresholds
in the (local as well as macroscopic) potential gradient. Then, depending on
the physics of the phenomenon under study, one of the following two scenarios
may arise.
26 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
(i) The transport properties of the material vanish below the threshold, but
above the threshold the material behaves linearly (or, possibly, exhibits
constitutive nonlinearity) and possesses non-zero effective transport prop-
erties. For example, consider a resistor network in which each bond is
insulating if the voltage drop between its two ends is less than a threshold
value, but becomes conducting (either linearly or nonlinearly) if the volt-
age drop exceeds a threshold. An example of a material to which such a
model is directly relevant is foam. As described in Chapter 9 of Volume I,
foams behave both as solid materials (in the sense of exhibiting an elas-
tic response when exposed to an external stress or strain), and as a fluid
when the applied stress that they are exposed to reaches a threshold value.
Therefore, foams do not flow if the stress applied to them is less than the
threshold. As a result, if we consider, for example, flow of foams in a
porous medium (which is usually modeled as a network of tubes), there
would be no macroscopic flux of foams unless the pressure gradient ap-
plied to the porous medium exceeds a threshold. We must, however, point
out that this type of threshold behavior is not the same as that of a percola-
tion system below and above the percolation threshold, i.e., this threshold
behavior is not a geometrical effect, although, as we will show in Chapter
3, there are certain similarities between the two types of phenomena.
(ii) The second scenario arises when the material behaves linearly (or, pos-
sibly, exhibits constitutive nonlinearity) if the applied potential gradient
is less than a threshold, but exhibits highly nonlinear properties when the
threshold is exceeded. Well-known examples of this type of phenomenon
are brittle fracture and dielectric breakdown of solid materials, phenomena
that will be studied beginning with Chapter 5.
Compared to linear systems, the number of studies in which an attempt has been
made to obtain estimates of the effective nonlinear properties is small. This is par-
ticularly true in the context of continuum models of disordered materials. Discrete
models have received much more attention, and will be described and discussed
in Chapter 3. To our knowledge, Marcellini (1978) was perhaps the first to un-
dertake a systematic study of effective transport properties of nonlinear materials,
and attempted to estimate their effective dielectric constant. He considered a two-
phase composite in which one phase had a constant dielectric constant, while the
dielectric constant of the second phase, that consisted of spherical inclusions, was
a function of the local electric field. The particles were arranged either at random
or in a periodic manner, similar to the periodic models that were described and
analyzed in detail in Chapter 4 of Volume I. Miksis (1983) obtained slightly more
general results for the effective properties of periodic arrays, and random distribu-
tions of nonlinear spherical inclusions in a linear matrix. The methods of Marcellini
and Miskis were more or less straightforward generalization of those described
in Chapter 4 of Volume I, and hence need not be described again. Willis (1986)
applied the approach of Hill (1963) (see below; see also Chapter 7 of Volume I
for more details) to nonlinear dielectrics. In terms of deriving rigorous bounds for
2.1. Variational Principles 27
the effective nonlinear electrical conductivity and dielectric constant, Talbot and
Willis (1985, 1987, 1994) and Willis (1986) proposed extensions of the Hashin–
Shtrikman variational principles (Hashin and Shtrikman, 1962a,b, 1963) (see also
Chapters 4 and 7 of Volume I) to nonlinear heterogeneous materials. In a series
of papers, Ponte Castañeda and co-workers (Ponte Castañeda, 1992b, 1998; Ponte
Castañeda and Kailasam, 1997) analyzed the effective nonlinear conductivity and
dielectric constant of two-phase heterogeneous materials using two different tech-
niques. One of the methods is exact to first-order in contrast between the properties
of the two phases, and is capable of delivering rigorous lower bounds and approx-
imate estimates for the upper bounds (not the upper bounds themselves), while
the second method is exact to second order in the contrast between the phases’
properties. To our knowledge, their work is the most advanced attempt in the area
of continuum description of the effective nonlinear conductivity and dielectric
constant of disordered materials, and is described in detail in this chapter.
where · · · denotes an spatial average. We should keep in mind that the effective
behavior of the heterogeneous material, as characterized by the energy functional
He (E), may in general be anisotropic, even if the material’s phases themselves
are isotropic. In principle, He (E) is determined by solving the usual electro-
static problem on , defined by, ∇ × E = 0, and ∇ · D = 0, subject to a uniform
boundary condition, ϕ = −E · x on the external surface of , where ϕ is the
electrostatic potential defined by, E = −∇ϕ(x) in . This boundary condition
ensures that the average of the electric field is in fact E, in the sense that
E = E(x) dx. (3)
Moreover, the average displacement field is defined by a similar relation:
D = D(x) dx, (4)
so that one obtains the effective energy He that evaluates the pertinent energy
functional for the heterogeneous material,
He (E) = w(x, E) dx, (5)
at the actual electric field solving the electrostatic problem for a given microstruc-
ture. Due to the complexity of the morphology of real materials, it is not practical
to solve the electrostatic problem. For this reason, variational formulations of
the problem based on the minimum energy and minimum complementary-energy
principles provide useful alternative routes for analyzing the problem. Thus, let
us state these principles here (which were also utilized in Volume I for obtaining
estimates of effective linear properties).
According to the minimum energy principle, expressed in terms of the energy
functional H, one can obtain the following expression for the effective energy He
of a heterogeneous material,
He (E) = min H(E), (6)
E∈S1
where
S1 = {E|E = −∇ϕ(x) in , and ϕ = −E · x on ∂}. (7)
Note that, to guarantee the existence of the minimizer (6), certain conditions on the
behavior of w (or e) as E → ∞ are required, which is why one assumes that w is
coercive. Moreover, strict convexity of He guarantees uniqueness of the solution,
convexity of w ensures that of He , and if the fields are smooth enough, Eq. (6)
will be equivalent to the original electrostatic problem defined above.
The second characteristic of the heterogeneous material is obtained from its
complementary-energy function Hec , defined in terms of the principle of minimum
complementary energy:
Hec (D) = min Hc (D), (8)
D∈S2
2.1. Variational Principles 29
where
Hc (D) = w ∗ (x, D) dx (9)
is the complementary energy functional, expressed in terms of
w∗ (x, D) = sup{E · D − w(x, E)}, (10)
E
with
S2 = {D|∇ · D = 0 in , and D · n = D · n on ∂} (11)
being the set of admissible electric displacement fields. Note that, if Eq. (3) is
reinterpreted as a definition for the average electric field, then, one has
∂
E = Hc (D). (12)
∂D e
In general, it can be shown that
Hec (E) ≥ Hec∗ (E). (13)
The reason for the inequality (13) is related to the fact that definitions of H and
Hc correspond to different boundary conditions on the heterogeneous material
(Dirichlet versus Neumann conditions), hence leading to generally distinct effec-
tive energies. However, the strict equality holds in (13) if the composite can be
homogenized, in the sense that it can be considered as homogeneous on a large
enough scale. Finally, note that
w ∗ (x, D) = e∗ (x, D), (14)
where e∗ is the convex polar function (Legendre transform) of e and D is the
magnitude of D.
Ponte Castañeda (1992b) proposed new variational principles in order to obtain
upper and lower bounds and estimates for the effective energy functions of non-
linear materials. These variational principles are equivalent to the standard ones
described above, under appropriate hypothesis on the energy-density function.
The new variational principle is based on a change of variables r = h(E), with
h : R + → R + (R + is the set of non-negative reals) given by h(E) = E 2 . One
than obtains a function f : × R + → R + , such that
f (x, r) = e(x, E) = w(x, E), (15)
has the same dependence on x as e and w, and that it is continuous and coercive (but
not necessarily convex) in r. Moreover, f is a non-negative function satisfying,
f (x, 0) = 0. Then, if we define the Legendre transform (convex polar) of f by
f ∗ (x, p) = sup{rp − f (x, r)}, (16)
r≥0
it follows that
f (x, r) ≥ sup{rp − f ∗ (x, p)}. (17)
p≥0
30 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
Note that x is fixed in (16) and (17), and that the suprema are evaluated over the
sets of non-negative r and p, respectively. In addition, the right-hand side of (17)
is the bipolar of f , which has the geometric interpretation of the convex envelope
of f , and hence the inequality. The equality in (17) is achieved if f is convex and
continuous in r. Therefore, assuming that the energy function w in (16) is such
that f is convex (note that convexity of f implies that of w), one obtains from (17)
the following representation for the local energy density function of the nonlinear
heterogeneous material,
w(x, E) = sup {w 0 (x, E) − v(x, 0 )}, (18)
0 ≥0
where p has been identified with 12 0 and r with E 2 , in such a fashion that
w0 (x, E) = 12 0 (x)E 2 and v(x, 12 0 ) = f ∗ (x, 0 ). Thus, w0 corresponds to the
local energy-density function of a linear, heterogeneous comparison material with
arbitrary (but not necessarily constant) non-negative dielectric constant 0 (x). The
minimum energy formulation of the variational principle follows by making use
of the representation (18) in the classical minimum energy principle, and inter-
changing the order of the infimum in (16) and the supremum in (18). The result
(Ponte Castañeda, 1992b) is the following theorem.
Theorem 1: Suppose that the local energy-density function w of a given nonlin-
ear heterogeneous material with isotropic phases satisfies condition (15) with
f a non-negative, continuous, coercive and convex function of r = E 2 , with
f (x, 0) = 0. Then, the effective energy function of the nonlinear heterogeneous
material He is determined by the variational principle,
He = sup {He0 (E) − V ( 0 )}, (20)
0 (x)≥0
where
V ( ) =
0
v[x, 0 (x)] dx, (21)
and He0 is the effective energy function of a linear heterogeneous comparison
material with local energy function w 0 , such that
He0 = min w 0 (x, E) dx. (22)
E∈S1
it follows that
g(x, s) ≤ inf {sq − g∗ (x, q)}, (25)
q≥0
with the equality holding true if g is concave. Assuming then that the complemen-
tary energy density function w ∗ of the nonlinear heterogeneous material is such
that g is concave, it follows from (25) that
w∗ (x, D) = inf {w 0∗ (x, D) + v(x, 0 )}, (26)
0 ≥0
where q has been identified with (2 0 )−1 and s with D 2 , such that w 0∗ (x, D) =
[ 12 0 (x)]D 2 is the complementary-energy function of the linear, heterogeneous
comparison material with arbitrary non-negative dielectric coefficient 0 (x), and
v(x, 0 ) = g ∗ (x, 12 0 ). Given these, one can state the following theorem (Ponte
Castañeda, 1992b)
Theorem 2: Suppose that the (convex) local complementary-energy function
w∗ of a given nonlinear heterogeneous material with isotropic phases satisfies
condition (23) with g being a non-negative, continuous, coercive and concave
function of s = D 2 , and g(x, 0) = 0. Then, the effective complementary-energy
function Hec of the nonlinear heterogeneous material is given by
Hec (D) = inf {He0c (D) + V ( 0 )}, (27)
0 (x)≥0
where
He0c (D) = min w 0∗ (x, D) dx (28)
D∈S2
where He0 is the effective energy function of a linear comparison material with N
phases of dielectric constants i0 with volume fractions φi , such that the effective
dielectric constant e0 of the comparison composite is given by
N
e0 (x) = mi (x)i0 . (32)
i=1
The function vi is given by Eq. (19), written for the ith phase, and the supremum
in (31) is evaluated over the set of constants i0 (i = 1, · · · , N).
Corollary 2: Suppose that the appropriate complementary version of (29)
characterizes the local complementary energy function w∗ of a N-phase non-
2.2. Bounds on the Effective Energy Function 33
where He0 = 12 ( 0e E) · E is the effective energy function of the linear material
with effective dielectric tensor 0e . The nonlinear Wiener lower bound for the
effective energy functions He of the nonlinear materials is obtained by applying
Eq. (31) to the set of nonlinear composites with given phase volume fractions,
and combining the result with the lower bound (35) for the corresponding linear
comparison materials. The result is
⎧ −1 ⎫
⎨1 N
φi N ⎬
He ≥ sup E 2
− φ v (
i i i
0
) , (36)
⎩
0 >0 2 i0 ⎭
i i=1 i=1
34 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
with
1 0 2
vi (i0 ) = sup i s − ei (s) . (37)
s>0 2
Clearly, the number of optimizations implicit in (36) and (37) is 2N , but this
number may be significantly reduced by using the identity,
N −1 N
φi
0
= inf φi i (1 − ωi ) ,
2
(38)
ωi
i=1 i i=1
where the infimum is over the set of variables ωi (i = 1, · · · , N) which are sub-
ject to a zero-average constraint, i.e., ω = N i=1 φi ωi = 0. This identity, when
applied to the nonlinear lower bound for He in (36), yields
N
1
He ≥ sup inf φi i (1 − ωi ) E − vi (i )
2 2 0
, (39)
0 >0
ωi 2
i i=1
In (40), the saddle point theorem and the fact that the argument of the nested
supremum and infimum is concave in i0 (since the functions vi are convex in
i0 ) and convex in ωi have been used in order to justify the interchange of the
supremum and infimum operations. Finally, application of Eq. (18), specialized to
each phase in the form
1 0 2
ei (s) = sup i s − vi (i0 ) , (41)
0 >0 2 i
leads to
He ≥ inf {φi ei (|1 − ωi |E)} , (42)
ωi
which is much simpler than the bounds (36) and (37), as it involves only a
N-dimensional optimization, with one linear constraint, which can easily be em-
bedded in the optimization operation by suitable relabelling of the optimization
variables. For example, for a two-phase material, bound (42) becomes
He ≥ inf {φ1 e1 (|1 − φ2 ω|E) + φ2 e2 (|1 + φ1 ω|E)}, (43)
ω
where (l) = inf s {s0 }. Equation (44), which is subject to the constraint that, ω =
0, may be rewritten as
N
e = inf
(l)
φi i (1 − ωi ) + (d − 1) ωi
2 (l) 2
. (45)
ωi
i=1
(46)
where the saddle point theorem has been used to justify interchanging the
supremum and infimum operations. Then, using (41), one obtains
⎧ ⎧
⎨ ⎨ N
He ≥ min inf φi ei (|1 − ωi |E)
s ⎩ ωi ⎩
i=1,i=s
⎫⎫
⎬⎬
1
N
+ φs es (1 − ωs )2 + (d − 1) φj ωj2 E , (47)
φs ⎭⎭
j =1
which represents the Hashin–Shtrikman lower bound for nonlinear isotropic ma-
(l)
terials with isotropic phases of given volume fractions, and is denoted by HHS .
36 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
either case), which is completely analogous to the corresponding result for linear
two-phase materials.
which must be optimized over the two scalar variables ω ∈ (−∞, ∞) and γ ∈
(−∞, ∞). The optimization can be carried out either analytically, if the energy
functions of the nonlinear phases are sufficiently simple, or numerically. Here, B
is a parameter which is given by (Torquato, 1985a,b)
(d − 1) − ζ2
B = (d − 1) . (53)
1 − (d − 1)ζ2
We can now consider two important limiting cases.
(n)
is defined by, He (E) = ge En /n, one obtains
n/2
(n)
ge ζ1 + (d − 1)φ2 + Bφ2 ζ2
= φ1 . (57)
(1)
g1 ζ1 φ12
Equations (56) and (57) can now be utilized for estimating the effective energy
of nonlinear composites with superconducting inclusions, provided that the ap-
propriate expressions for the microstructural parameters ζ1 and ζ2 = 1 − ζ1 are
available, a matter that was discussed in detail in Section 4.5.3 of Volume I. In
particular, it can be shown that Eq. (56) provides an estimate of the effective energy
which is always larger than the estimates provided by Eqs. (48) and (51), hence
satisfying these rigorous bounds.
(u)
N
HW ≤ φi ei (E). (63)
i=1
The determination of an estimate for the Hashin–Shtrikman upper bound, or
the upper estimate, is accomplished by application of approximation (31) to the
Hashin–Shtrikman upper bounds for the linear comparison material. The upper
bound for the effective energy function of the linear comparison material may be
given in terms of the upper bound for its effective dielectric constant:
N −1
φ
e+ = − (d − 1) + ,
i
0 + (d − 1) +
(64)
i=1
i
where + = supi {i0 }. The procedure that utilizes the lower bound (44) for the
linear comparison material to obtain a lower bound for the nonlinear material may
now be repeated. To derive the upper estimates, one utilizes (64) instead of (44),
in which case the result would be the same as (47) and (48) for the N -phase and
two-phase nonlinear materials, respectively, with the difference that the outermost
minimum operations must now be replaced by maximum operations. The result,
+
denoted by HHS , is referred to as the Hashin–Shtrikman upper estimate. However,
+
as shown below, HHS is not, in general, an upper bound for He .
The same arguments and analyses also apply to the Beran upper bounds. Hence,
one can obtain upper estimate for the Beran-type bounds. If
2
+
= ζi i0 , (65)
i=1
then, the corresponding result for the upper estimate (which, in general, is not an
upper bound) for nonlinear isotropic materials is given by
"
+
HB (E) = inf φ1 e1 (1 − φ2 ω)2 + (d − 1)φ2 ζ1 ω2 E
ω
"
+ φ2 e2 (1 + φ1 ω)2 + (d − 1)φ1 ζ2 ω2 E . (66)
40 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
linear laminate lies within and attains (for specific orientations of the applied fields)
the Wiener bounds. Thus, at least in this case, the Wiener bounds on the effective
energy function of arbitrarily anisotropic-linear materials are sharp.
In the context of two-phase linear materials, it is known that only iterated lami-
nates of rank greater than or equal to the dimension of the underlying physical space
(d = 2 or 3) can have isotropic properties. The isotropy is obtained by choosing
the relative volume fractions and the layering directions of each of the embedded
laminates in such a way that the tensor representing the effective property of in-
terest is isotropic, while the absolute volume fractions of the constituent phases
remain fixed. One might criticize sequentially-laminated materials by noting that
the inclusions are flat, whereas in practice the inclusions are often equi-axed.
However, one must note that iterated laminates can be used to model arbitrarily
close the properties of any two-phase microstructure (Milton, 1986). For exam-
ple, the coated-spheres model of Hashin and Shtrikman (1962a,b, 1963) possesses
exactly the same effective properties as an isotropic iterated laminate with the
same volume fractions. In the coated-spheres model (see also Chapters 3, 4 and
7 of Volume I) the material consists of composite spheres that are composed of a
spherical core of conductivity g2 and radius a, surrounded by a concentric shell of
conductivity g1 with an outer radius b > a. The ratio a/b is fixed, and the volume
fraction φ2 of inclusions in d dimensions is given by φ2 = (a/b)d . The composite
spheres fill the space, implying that there is a sphere size distribution that extends
to infinitesimally-small spheres. In Chapters 4 and 7 of Volume I we derived exact
expression for the effective conductivity and elastic moduli of the coated-spheres
model and low-rank laminates.
As shown in Chapters 4 and 7 of Volume I, the Hashin–Shtrikman bounds for
the coated-spheres model, which represent isotropic microstructures, are exact es-
timates. Thus, it may seem that the coated-spheres model may be more realistic
than the iterated laminates. However, the laminates have a distinct advantage over
the coated-spheres model in that, they contain a finite number of length scales, in
contrast with the coated-spheres microstructure which involves an infinite number
of length scales because, as described above, the composite spheres must cover
all sizes to fill the space. Another advantage of sequentially-laminated materials
is that, when subjected to uniform boundary conditions, the fields are piecewise
constant within the material (regardless of whether the composite’s phases are
linear or nonlinear), except in small boundary layer regions at the interfaces sep-
arating laminates of different ranks, the effect of which is made negligible by the
hypothesis of separated length scales. This fact was used for deriving the exact
results for the effective linear properties of the laminates presented in Chapters 4
and 7 of Volume I.
To compute the effective energy function of nonlinear rank−d laminates (d = 2
and 3) with layering directions n1 , · · · , nd , we denote by φI the volume fraction of
phase 1 with energy function e1 in the first-rank laminate, and note that 1 − φI is the
corresponding volume fraction of phase 2 with energy e2 . Ponte Castañeda (1992b)
showed that the effective energy function of the nonlinear, first-rank laminate is
42 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
given by
HeI (E) = inf {φI e1 (s1 ) + (1 − φI )e2 (s2 )} , (67)
(1) (2)
ωI ,ωI
(1) (2)
subject to the constraints that ωI = φI ωI + (1 − φI )ωI = 0, and where
"
(1)
s1 = E2 − E12 + [1 − ωI ]2 E12 ,
"
(2)
s2 = E2 − E12 + [1 − ωI ]2 E12 , (68)
where E1 = E · n1 .
Consider now the second-rank laminate obtained by mixing layers of the first-
rank laminate with layers of a third phase characterized by an energy function
e3 and relative (to the second-rank laminate) volume fractions φI I and 1 − φI I ,
respectively. The new lamination direction n2 is orthogonal to n1 . Then, the fol-
lowing energy function for the nonlinear second-rank laminate in dimension d ≥ 2
is obtained (Ponte Castañeda, 1992b):
HeI I (E) = inf {φI I φI e1 (s1 ) + φI I (1 − φI )e2 (s2 ) + (1 − φI I )e3 (s3 )} ,
(1) (2) (1) (2)
ωI ,ωI ,ωI I ,ωI I
(69)
subject to the constraints that ωI = ωI I = 0 (where ωI I is defined in a
manner analogous to ωI ), and where
"
(1) (1)
s1 = E2 − E12 − E22 + [1 − ωI ]2 E12 + [1 − ωI I ]2 E22 ,
"
(2) (1)
s2 = E2 − E12 − E22 + [1 − ωI ]2 E12 + [1 − ωI I ]2 E22 , (70)
"
(2)
s3 = E2 − E22 + [1 − ωI I ]2 E22 ,
where Ei = E · ni .
Asimilar result can be obtained for a two-phase, nonlinear second-rank laminate.
In this case, the result for HeI I is generally anisotropic and direction-dependent,
but it may be used in two dimensions (2D) for deriving an isotropic result, for
each value of E, by an appropriate choice of φI I (but not the choice that makes
the corresponding linear second-rank laminate isotropic), which is obtained by
requiring that φI I (0 ≤ φI I ≤ 1) and E1 satisfy the following relations
∂HeI I ∂HeI I
= 0 and = 0, (71)
∂E1 ∂φI I
where the first relation is subject to the constraint that, E12 + E22 = E2 , while
in the second relation one assumes that E is fixed. These conditions follow by
performing a Taylor series expansion of (69) in φI I and E1 and requiring that
the expansion yield the same result for any choice of φI I and E1 . Physically, this
corresponds to selecting a microstructure (by choosing φI I )—with fixed over-
all volume fractions of the phases—for each value of E, ensuring that HeI I is
2.4. Effective Dielectric Constant of Strongly Nonlinear Materials 43
independent of the direction of E, thus guaranteeing that the resulting energy
function is isotropic. However, the resulting energy function does not correspond
to a fixed microstructure, rather to a family of (anisotropic) microstructures, each
one of which is obtained from one value of the applied electric field.
The effective energy function of a nonlinear third-rank laminate is obtained by
analyzing the effective behavior of a simple laminate made up of layers of the
second-rank laminate and of layers of a fourth phase with energy function e4 and
volume fractions φI I I and 1 − φI I I , respectively. The new layering direction n3
is selected to be orthogonal to both n2 and n1 . Then, the effective energy function
of the nonlinear third-rank laminate is given by (Ponte Castañeda, 1992b)
HeI I I (E) = inf {φI I I φI I φI e1 (s1 )
(1) (2) (1) (2) (1) (2)
ωI ,ωI ,ωI I ,ωI I ,ωI I I ,ωI I I
the laminates, and both differ strongly from the Wiener bound, with the latter
yielding estimates that are larger than the former two.
(83)
where p is the root of the quadratic equation,
11−y 1
(2 − x)p 2 − xy + 2(1 − y) p + (1 − x)(1 − y) = 0,
21−x 2
and
xy 1 − x
q= .
xy − φ 2 − x
In this case, the Hashin–Shtrikman upper estimates for the isotropic composite
lie well below the Wiener bounds for arbitrarily anisotropic composites. On the
other hand, the exact estimates for the nonlinear isotropic laminates lie above the
Hashin–Shtrikman upper estimates, hence verifying that the Hashin–Shtrikman
upper estimates are not in general upper bounds. This is due to the fact that
the isotropic laminates correspond to specific microstructures within the class
of isotropic composite materials, and if the Hashin–Shtrikman upper estimates
were rigorous bounds for such materials, they would have to lie above all possible
isotropic microstructures, and, in particular, they must lie above the isotropic lam-
inates. Nevertheless, the effective dielectric constants of the isotropic laminates
are not far from the Hashin–Shtrikman upper estimates.
(n)
where gi is the generalized nonlinear conductivity of phase i. The linear
comparison materials are now defined by the quadratic energy-density function,
1 0
w 0 (x, E) = g (x)E 2 , (85)
2
where g 0 (x) is the conductivity of the fictitious linear material. Then, under the
hypothesis that the functions ei of the nonlinear material are convex on E 2 , the
analogues of Eqs. (37) and (41) for the conductivity problem are given by
1 0
ei (E) = max g (x)E 2 − vi (g 0 ) , (86)
g 0 ≥0 2
1 0 2
vi (g 0 ) = max g E − ei (E) . (87)
E 2
Note that if the functions ei are smooth, the maxima in Eqs. (86) and (87) are
attained at
1 2 ∂vi 1 ∂ei
E = 0 , g0 = , (88)
2 ∂g E ∂E
respectively, which are inverse of each other. Then, the analogue of Eq. (31) for
the conductivity problem is given by
N
He (E) = max He (E) −
0
φi vi [g (x)]i ,
0
(89)
g 0 (x)≥0
i=1
where He0 is the effective energy function of the linear comparison material, with
local energy function (85), such that
He0 (E) = min w 0 (x, E), (90)
E∈S2
constant over the individual phases, unless the actual fields are constant through-
out the phases. However, as discussed earlier in this chapter, a lower bound for
He0 can be obtained by restricting the class of trial comparison conductivity fields
to be constant within each phase such that
N
g 0 (x) = mi (x)gi0 , (91)
i=1
where gi0 is constant, and mi (x) is the indicator function of phase i, as before.
Equation (91) follows from the fact that the maximum over a set is, in general,
larger than the maximum over any subset of the original set. Therefore, from
Eqs. (89) and (91), it follows that
N
He (E) ≥ max He0 (E) − φi vi (gi0 ) , (92)
gi0 >0 i=1
where
1
N
1
He0 (E) = E · [ge(l) E] = min φi gi E i .
0 2
(93)
2 E∈S2 2
i=1
(l)
Here, ge is the effective conductivity tensor of a linear comparison material with
precisely the same morphology as the original nonlinear composite which, in
general, is anisotropic.
As discussed earlier in this chapter, the above estimates for N -phase nonlinear
(l)
materials represent lower bounds for He . Thus, lower bounds for ge may be
(l)
used to generate lower bounds for He , but upper bounds for ge cannot be used
for deriving upper bounds for He . In this case, one can ignore the inequality in
(92) and reinterpret it as an approximate equality in order to obtain estimates for
specific types of materials. Denoting by ĝi0 the optimal values of gi0 from Eq. (92),
it follows that the average current field I is given by
N
1
(l)
∂ge ∂vi 0 ∂ ĝi0
I = ge(l) (ĝ10 , · · · , ĝN
0
)E + E · ( ĝ 0
1 , · · · , ĝN
0
)E − φ i ( ĝ i ) ,
i=1
2 ∂gi0 ∂gi0 ∂E
(94)
so that, the maximum in (92) for the general bound is attained at
(l)
1 ∂ge ∂vi
E · 0
(ĝ10 , · · · , ĝN
0
)E = φi 0 (ĝi0 ) (i = 1, · · · , N). (95)
2 ∂gi ∂gi
The constitutive relation that defines the effective conductivity of the nonlinear
material reduces to
I = ge(l) (ĝ10 , · · · , ĝN
0
)E. (96)
Note that, Eq. (96) is fully nonlinear because the variables depend nonlinearly ĝi0
(l)
on E. Since the linear conductivity ge is a homogeneous function of degree
48 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
one in the conductivity constants gi0 of the linear comparison material (see also
Chapters 2, 4 and 6 of Volume I), then
N (l)
∂ge
gi0 = ge(l) . (97)
i=1
∂gi0
where
1/2
(l)
1 ∂ge
Êi = E · (ĝ1 , · · · , ĝi )E
0 0
(i = 1, · · · , N). (100)
φi ∂gi0
Finally, the constitutive relation describing the effective behavior of the nonlinear
material is written in the following form,
1 ∂e1 1 ∂eN
I = ge(l) (Ê1 ), · · · , (ÊN ) E, (101)
Ê1 ∂E Ên ∂E
where Êi are functions of the (average) applied field E, the nonlinear properties
of the constituent phases of the material, and the material’s microstructure.
and therefore
1 1 ∂e1 1 ∂eN
He (E) ≥ E · ge
(l)
(Ê1 ), · · · , (ÊN ) E . (104)
n+1 Ê1 ∂E ÊN ∂E
2.5. Effective Conductivity of Nonlinear Materials 49
Figure 2.2. Comparison of various bounds and estimates for the effective nonlinear conduc-
(n)
tivity ge of 2D, isotropic, two-phase, power-law conductors with power exponent n = 3
(n) (n)
and g2 /g1 = 1000. Symbols show the results of numerical simulations with random
resistor networks (RRN). The lower bound MG(l) obtained from the Maxwell–Garnett ap-
proximation is identical with that obtained from the Hashin–Shtrikman lower bound. MG(u)
and B(u) denote, respectively, the estimates for the upper bound using the Maxwell–Garnett
approximation and the Beran upper bound. (after Ponte Castañeda, 1998).
7 of Volume I), the Beran bounds provide a way of estimating the effective proper-
ties of more general types of microstructures for which the Maxwell–Garnett and
EMA estimates may not be accurate.
Equation (111) assumes that the reference fields E(i) are known in terms of the
λ(i) , which, in general, are functions of the actual electric field E, as well as of
the (unknown) Ei , and therefore the problem posed by (111) for He is nonlinear.
However, provided that the second derivatives of wi in Eq. (107) vary slowly
with the E(i) , Eq. (111) suggests, as an approximation, replacing E(i) by a (as-yet
unknown) constant, in which case ĝ(i) , Pi and vi will also be constant within each
phase, hence leading to the following expression for He ,
N
He (E) = φi vi (E) + P̃ (E), (113)
i=1
where
1
P̃ (E) = min P · E + E · (ĝE ). (114)
E ∈S 2
The interesting feature of Eq. (113) is that it requires only the solution of the
linear problem (114) for P̃ which is, physically, equivalent to a problem for a
linear conductor with N anisotropic constituents with conductivity tensors ĝ(i) and
prescribed polarizations Pi , a problem much simpler to analyze than the original
nonlinear problem for He . The question then arises as to what the best choices are
for these constants.
The optimal choice for each Ei is Ei , the average of the actual field E over
phase i:
Ei = Ei , (115)
where · · ·i denotes a volume average over phase i. Although Ei cannot be
obtained exactly, a consistent estimate for it may be obtained by noting that, Ei =
E + E i , where
1 ∂ P̃
E i = (116)
φi ∂Pi
with ĝ(i) held fixed. On the other hand, although the best choice for the E(i) is not
a priori clear, given the approximation that was made in deriving (113), the choice
E(i) = Ei , (117)
2.6. Second-Order Exact Results 53
is simple and plausible. In particular, Eqs. (115) and (117) are exact for laminated
materials, where the fields are constant within each phase. Thus, any solution for
the problem posed by (114), together with the associated estimates (116), may
be utilized for obtaining corresponding estimates for He via Eq. (113), together
with the (self-consistent) equations (115) and (117). Note that E(i) = Ei , and,
for this reason, ĝ(i) is henceforth denoted by gi , the phase conductivity tensor. In
particular, for two-phase composite materials one can show that
1 (l)
P̃ (E) = (ge − g)(g)−1 P · (g)−1 P, (118)
2
from which it follows, using (116), that
1
E1 = E + (g)−1 (ge(l) − g)(g)−1 P, (119)
φ1
1
E2 = E + (g−1 )(ge(l) − g)(g)−1 P, (120)
φ2
(l)
where g = g1 − g2 , P = P1 − P2 , g = φ1 g1 + φ2 g2 , and ge is the effective
conductivity tensor of a two-phase linear material with phase conductivity tensors
g1 and g2 , volume fractions φ1 and φ2 , and precisely the same microstructure as
the nonlinear composite. This means that any estimate that is available for the
(l)
effective conductivity tensor ge of a two-phase linear material, including, for
example, the Maxwell–Garnett and EMA estimates, can be used for generating
the corresponding estimates for He of a two-phase nonlinear material. Note that
the approximate estimate of He given by Eq. (113) is a convex function. Since
the exact expression for He is also known to be convex, it follows that deriva-
tives of the approximate expressions for He should provide a reasonably accurate
approximation to the exact constitutive relation.
with
(p)
gi = gi [1 + (1 − n)ωi ], (123)
(v) 1
gi = gi 1 + (1 − n)ωi + n(n − 1)ωi2 . (124)
2
Then, from Eqs. (84), (113) and (118) we obtain
n + 1 (l) g (p) − g (p) 2
ge(n) = g (v)
+ mge − ng 1 2
, (125)
2n2 g1 − g 2
(n) (n)
where here gi = gi |1 + ωi |n−1 , with gi being the nonlinear conductivity of
phase i, and g (v) defined in a manner analogous to g. The variables ωi are
determined by Eqs. (115), (119), (120) and (121); they yield, ω1 = φ2 ω, ω2 = φ1 ω,
where
(l) (p) (p)
1 1 mge − ng g1 − g2
ω= . (126)
φ1 φ 2 n 2 g1 − g 2 g1 − g 2
(n)
Estimates of ge based on the Maxwell–Garnett approximation and the EMA
can now be obtained by using their corresponding estimates for the effective con-
(l)
ductivity tensor ge , which are in terms of the phase conductivity tensor gi . Since
the Maxwell–Garnett approximation is not symmetric in material’s phases, one
obtains two classes of Maxwell–Garnett estimates, corresponding to particulate
microstructures with the less and more conducting material designated as the ma-
trix phase. On the other hand, due to its symmetry, the estimates provided by the
EMA are unique. Moreover, it should be pointed out that the Maxwell–Garnett
(n)
and EMA estimates for the effective nonlinear conductivity ge are not exactly
equivalent. In fact, it can be shown that while for sufficiently weak nonlinearity
(i.e., for n 1) these estimates are in close agreement with each other, they can be
significantly different for stronger nonlinearities (i.e., as n → 0 or ∞). The reason
for the differences is associated with the nature of the approximations made in
going from the exact estimate (111) for He to the approximation (113), assuming
that the reference conductivity tensors ĝ(i) vary slowly with E(i) , so that the re-
placement of the E(i) by Ei does not introduce significant errors. In what follows,
we summarize the results obtained with the Maxwell–Garnett and EMA estimates.
The details of derivation of these results, which is straightforward, are given by
Ponte Castañeda and Kailasam (1997).
and
g − (g1 + g2 )/β(m)
ge(l) =
2[1 − 1/β(m)]
#
g − m(g1 + g2 )/β(m) 2 g 1 g2
+ + (131)
2[1 − m/β(m)] β(m)/m − 1
where
⎧ √
⎨ 1+ n, 2D,
$ " %−1
β(n) = (132)
⎩ 2(1 − n) 1 − √n arcsin n−1
n , 3D,
n−1
and g = φ1 g1 + φ2 g2 , as before. Equations (130) and (131), obtained from the
(l)
two independent components of the anisotropic tensor ge , depend on the func-
tions α and β which, in turn, are known functions of the unknown parameter m.
(l)
Therefore, m is obtained by equating (130) and (131). Once m is obtained, ge
(n)
and hence ge are computed.
We now consider the application of these results to estimating the effective
nonlinear conductivity of two important classes of heterogeneous materials that we
have been studying throughout this book, namely, those with superconducting or
insulating inclusions.As we emphasized in Volume I, because these two composites
represent two extreme limits of contrast between the properties of the two phases,
they provide stringent tests of any theory. In other words, if a theory is reasonably
accurate in these limits, it will be even more accurate in less extreme cases.
56 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
with ω = [φ1 + α(n) − n]−1 . The corresponding EMA estimates are obtained
from Eq. (135) with ω = [φ1 + nα(m)φ2 /m − 1]−1 , where m is the root of the
following equation
φ1 β(m) φ1 α(m)
m 1+ =n 1+ . (136)
1 − β(m) m − α(m)
The EMA estimates are valid for φ1 ≤ 1 − β(m)−1 . The limit φ1 = 1 − β(m)−1
(n)
defines the percolation threshold at which ge vanishes.
2.6. Second-Order Exact Results 57
(n)
Figure 2.3. The effective resistivity Re of 3D, isotropic, two-phase, power-law materials,
as predicted by the various approximations, versus the volume fraction φ1 of the supercon-
ducting inclusions, with n = 3. Note that only the effective-medium approximation indi-
cates the existence of a percolation threshold (after Ponte Castañeda and Kailasam, 1997).
05
Figure 2.4. Same as in Figure 2.3, but with insulating inclusions (after Ponte Castañeda
and Kailasam, 1997).
with
1 n+1
γ (n) = −1 . (138)
n 2[α(n) − n]
Analogous expressions can also be derived for the Wiener and Hashin–Shtrikman
bounds. Figure 2.5 presents a comparison of γ (n) for the MG/EMA estimates ver-
sus the Wiener and Hashin–Shtrikman bounds for the 2D materials, along with the
numerical results of Lee and Mear (1992), who reported their results for transverse
shear of fiber-reinforced power-law ductile composite conductors. As one might
expect, the MG/EMA estimates lie above the rigorous Wiener lower bound (for
the resistivity) for particulate microstructures. Moreover, the new estimates are in
excellent agreement with the numerical estimates of Lee and Mear (1992).
Figure 2.5. Dependence of the coefficient γ on the power-law exponent n, for a 2D,
isotropic, two-phase materials with insulating inclusions. Symbols represent the numerical
results of Lee and Mear (LM) (1992) (after Ponte Castañeda and Kailasam, 1997).
(n) (n)
for a given value of the ratio R1 /R2 , depending on whether phase 1 or 2 is
designated as the matrix phase (and vice versa for the inclusion phase), we restrict
(n) (n)
our attention to R1 /R2 > 1, and denote by MG1 and MG2 the two estimates
corresponding to designating phases 1 and 2, respectively, as the matrix phase.
Ponte Castañeda and Kailasam (1997) showed that as the volume fraction φ1
(n)
of the inclusions increases, the MG2 estimates for Re also increase. However,
the rate of the increase decreases with increasing n. In particular, for sufficiently
(n) (n)
large n, there is hardly any increase in Re over the matrix resistivity R2 . The
reason for this effect is the fact the current density becomes concentrated in the
more conducting matrix phase as n increases, and therefore the effect of the inclu-
sions becomes insignificant. Moreover, as the volume fraction φ1 of the inclusions
(n)
increases, the MG1 estimates for Re decrease, with the rate of the decrease in-
creasing with increasing n. In addition, as is the case for the estimates of the linear
(n)
EMA (see Chapter 4 of Volume I), the nonlinear EMA estimates for Re agree
with the MG1 and MG2 estimates in the limits of small volume fractions of phases
2 and 1, respectively.
(n)
The 3D Maxwell–Garnett and EMA estimates for Re also agree with the corre-
sponding small-contrast asymptotic results of Blumenfeld and Bergman (1991b),
which are known to be exact to second order in the contrast, and with the Wiener
upper and lower bounds (see above). In fact, the Maxwell–Garnett and EMA es-
60 2. Nonlinear Conductivity and Dielectric Constant: The Continuum Approach
(n) (n)
timates for ge and Re reproduce the asymptotic estimates of Blumenfeld and
Bergman:
n + 1 g 2 − g2
ge(n) ∼ g − , (139)
2α(n) g
n + 1 R 2 − R2
Re(n) ∼ R − , (140)
β(n) R
where R = φ1 R1 + φ2 R2 , with R1 and R2 being the resistivities of phases 1
and 2, respectively. The agreement for small enough contrast (to second-order)
is a consequence of the fact that the effective behavior of weakly heterogeneous,
nonlinear materials with statistically isotropic microstructures is dependent only
upon the phase volume fractions (Blumenfeld and Bergman 1991). However, while
(n)
the small-contrast expansions of Blumenfeld and Bergman (1991) for ge and
(n)
Re diverge as n → 0 and ∞, respectively, and can therefore yield unphysical
results even at relatively small contrasts, the estimates provided by Eq. (125) do
not diverge and always yield physically meaningful results. The new Maxwell–
Garnett estimates presented in this section satisfy all the known rigorous bounds,
including the Wiener bounds and the Hashin–Shtrikman upper bounds of Ponte
Castañeda (1992b) derived earlier in Sections 2.4 and 2.5, and the lower bounds of
Talbot and Willis (1994, 1996) for nonlinear composites with statistically isotropic
particulate microstructures (with n ≥ 1).
Finally, let us point out that Gibiansky and Torquato (1998b) derived cross-
property bounds that link the effective conductivity of nonlinear disordered
materials to their effective elastic moduli. Such cross-property bounds were already
described in Section 7.9 of Volume I for linear materials, and will be presented in
Chapter 4 for nonlinear composites.
Summary
In this chapter, we described and discussed general procedures for estimating
the effective conductivity and dielectric constant of nonlinear materials. These
procedures, which represent the generalization of those described in Chapters 4 and
7 of Volume I for linear materials, provide bounds and estimates for the effective
conductivity and dielectric constant. One procedure leads to rigorous bounds and
estimates that are exact to first order in the phase property contrast, while the
second technique yields estimates that are exact to second order in the contrast.
One important difference between the results obtained by the two procedures
must be emphasized. By design, the results presented in Section 10.6 are exact
to second-order in the phase contrast, and thus are consistent with the asymptotic
results of Blumenfeld and Bergman (1991b), whereas the results presented in Sec-
tions 10.4 and 10.5 are nonlinear estimates that are exact only to first order in the
phase contrast. On the other hand, while the first-order results provide rigorous
bounds for the effective energy function of nonlinear materials (and hence their
2.6. Second-Order Exact Results 61
3.0 Introduction
In this chapter we study nonlinear transport and optical properties of heteroge-
neous materials, representing their morphology by a discrete model. In particular,
we consider two-phase materials with percolation disorder which represents a
strong type of heterogeneity, although all the theoretical developments that are
described in this chapter (and throughout this book) are equally applicable to other
types of disorder. As we emphasized in Volume I, we believe that if a theory can
provide accurate predictions for transport and optical properties of materials with
percolation disorder, i.e., materials in which the contrast between the properties of
its two phases is strong, it should also be able to do so for almost any other type
of disorder.
There are many transport processes in which the current density is not related
to the applied field through a linear relation. Such nonlinearities, in the limit of
zero frequency, play an important role in many phenomena, including dielectric
breakdown, field dependence of hopping conductivity in heavily-doped semicon-
ductors, and many others. They are, at finite frequencies, the basis of nonlinear
optical phenomena in many disordered materials. By suitably tuning of the mate-
rial’s parameters, such as the volume fraction of the conducting material and its
nonlinear susceptibility, one can design a wide variety of composite materials with
specific properties that have important industrial applications. Chapter 2 described
the theoretical methods for estimating the effective conductivity and dielectric
constant of nonlinear disordered materials, based on the continuum models. In the
present chapter we consider several classes of nonlinear transport processes and
describe and discuss, based on the discrete models of heterogeneous materials, the
progress that has been made in understanding such phenomena.
Because Gkj has a lower bound, so does the function F , and therefore it has
a minimum. The existence of this minimum is equivalent to the existence of a
solution to Kirchhoff’s equations for the resistor network. This can be easily shown
by calculating ∂F /∂vk and showing that it vanishes at node k, hence demonstrating
that the net current reaching node k is zero.
However, because this is a nonlinear system, the proof is complete only one
also proves that, in addition to existing, the solution to Kirchhoff’s equations is
also unique. This can also be proven (Straley and Kenkel, 1984) by assuming that
(1) (2)
the function F has two minima for the voltage distributions {vk } and {vk }. If
(s)
so, then F must also have a saddle point at {vk }, because along any path in the
(1) (2)
voltage space that connects the two distributions {vk } and {vk }, F must have a
maximum. If this saddle point exists, it must be a solution to Kirchhoff’s equations,
∂F /∂vk = 0. However, if the function i(v) [e.g., one that is defined by Eq. (1)]
is differentiable with a positive derivative, then it is not difficult to show that the
64 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
k,j
In this equation, the linear term must vanish as ∂Gkj /∂vk = 0, and the quadratic
terms are all positive since we assumed that, di/dv > 0, and therefore the saddle
point does not exist, implying that the function F has a unique minimum, i.e., the
solution to Kirchhoff’s equations is unique. We note that Larson (1981) showed
that for slow flow of a power-law fluid in a porous medium with one injection
point and one producing point (which is the analogue of a two-terminal network)
in which flow in each pore is governed by Eq. (1), an equation similar to (1) is
also valid at the macroscopic scale, i.e., one has, at the macroscopic scale, Q =
GP 1/n , or, Q = G(P /L)1/n , where L is the length of the porous medium.
The reason that the general form of power-law (1) survives at the macroscopic
scale is that, such power-laws are self-similar and therefore they preserve their
identity under a microscopic-to-macroscopic transformation (that is, power laws
propagate self-similarly).
Calculating the voltage distribution in a nonlinear resistor network is a difficult
task, since the nonlinear Kirchhoff’s equations may have multiple solutions (all
but one of which would be unphysical), and thus one must be careful with the
numerical technique used in the simulation (see Uenoyama et al., 1992, for a
discussion of this point).
Suppose now that the bonds’ conductances are distributed according to a proba-
bility density distribution f (g). We derive an integral equation, from the solution
of which all the properties of interest can be computed (Sahimi, 1993a). Consider
3.1. Strongly Nonlinear Composites 65
For an infinitely large Bethe lattice, G and Gi are statistically equivalent. Thus, if
H (G) represents the statistical distribution of G, we must have
⎧ Z−1 $ ⎫
⎨ 1 %−1/n ⎬ Z−1
'
1
H (G) = · · · δ G − + n f (gi )H (Gi )dgi dGi .
⎩ gin Gi ⎭
i=1 i=1
(7)
If we now take the Laplace transform of both sides of Eq. (7), we obtain (Sahimi,
1993a)
∞
H̃ (s) = exp(−sG)H (G)dG
0
$ %−1/n Z−1
1 1
= exp −s + n f (g)H (G)dgdG . (8)
gn G
From the numerical solution of integral equation (8) we obtain all the properties of
interest. Note that, in the limit n = 1, Eq. (8) reduces to the corresponding integral
equation for Bethe lattices with linear resistors which was analyzed in Chapter
66 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
5 of Volume I. To our knowledge, no exact solution of Eqs. (8) and (9), for any
distribution f (g) and any value of n, has been derived.
(Straley, 1977). To estimate ge one may proceed as follows (Straley and Kenkel,
1984). The average power P dissipated per unit volume is given by
$ %1+1/n
IV V
P= = ge , (13)
AL L
where A and L are, respectively, the surface area and linear size of the sample.
The voltage difference across a chain of the lattice is controlled by the geometrical
distance ξp between the ends of the chain, where ξp is the correlation length of
percolation. In general, ξp is less than L, the length of the chain, since the chain is
twisted. However, in a Bethe lattice, the chain performs a random walk in space,
implying that, ξp2 ∼ L, and therefore the current Ic that is carried by a chain is
given by
$ % $ %1/n
ξp V 1/n 1/2n V
Ic = = (p − pc ) . (14)
LL L
However, the chain will carry no current at all unless its ends are connected to the
sample-spanning percolation cluster. To calculate the probability of this connec-
tion, we note that the probability that a given site is connected by a particular bond
to the sample-spanning cluster is P (p), the percolation probability, and therefore,
near pc , the two ends of the chain are connected to the cluster with a probability
P 2 (p) ∼ (p − pc )2β . As β = 1 for the Bethe lattice, we find that the probabil-
ity that the chain is connected to the sample-spanning cluster is proportional to
(p − pc )2 . Therefore, the dissipated power is given by
P = P 2 (p)[(p − pc )1/2n (V /L)1/n ]1+n = (p − pc )(5+1/n) (V /L)1+1/n , (15)
which, when compared with Eq. (13), implies that, near pc ,
ge ∼ (p − pc )(5+1/n)/2 . (16)
Observe that the critical exponents that characterize the near threshold behavior
of both gm and ge depend on n. In particular, Eq. (16) indicates that if, in general,
near pc one has
ge ∼ (p − pc )µ(n) , (17)
where µ(n) is the analogue of the conductivity critical exponent µ for the linear
case; that is, for linear resistor networks near pc one has
ge ∼ (p − pc )µ , (18)
then in the mean-field approximation (the solution of which is obtained by solving
the problem on a Bethe lattice) µ(n) = µn is given by
1
µn = (5 + n−1 ), (19)
2
which implies that, in the linear (n = 1) limit, one has
ge ∼ (p − pc )3 , (20)
in agreement with the result derived in Chapter 5 of Volume I.
68 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
One of the first EMAs for resistor networks with power-law conductances was
proposed by Sahimi (1993a), and is given by
∞
gZ/2
− 1 f (g)dg = 0, (25)
0 [g n + ((Z/2)n − 1)Gn ]1/n
which reduces to Eq. (24) in the limit n = 1, as it should. Another EMA was de-
rived by Tua and Bernasconi (1988) for a 2D isotropic continuum (with circular
inclusions), which was extended (Sahimi, 1993a) to networks of random conduc-
tances with coordination number Z. In this approach one first defines a tangent or
differential conductance σ by
di
σ = , (26)
dv
which, in the limit n = 1, yields the usual σ = g. Equation (26), when combined
with (1), yields
g
σ = v (1−n)/n . (27)
n
Consider now a two-phase material with its phase tangent conductances being σ1
and σ2 , both of which depend on the voltage v. Recall from Chapter 5 of Volume I
that in the EMA approach one inserts in the effective medium a bond with its true
conductance and determines the voltage fluctuations along this bond, i.e., the extra
voltage in the effective medium generated by the replacement of the conductance
of the bond in the effective medium by its true value. Carrying out this replacement
for component j (j = 1, 2) yields
σe (v1 , v2 )Z/2
vj = ve , (28)
σj (vj ) + σe (Z/2 − 1)
where ve is the voltage along the bond in the effective medium, and σe is the
effective value of σ . If we now apply the usual idea of an EMA, namely, that the
average of vj must be equal to ve (or that the average of the voltage fluctuations
must be zero), we obtain
∞
σj (vj ) − σe
f (σj )dσj = 0, (29)
0 σj + σe (Z/2 − 1)
which is the same as Eq. (24) except that the conductances σj and σe are functions
of the voltage. If the composite consists of two phases with (volume) fractions p
and (1 − p), then
pv1 + (1 − p)v2 = ve . (30)
The generalization of Eq. (30) to an N-component system is obvious. Equations
(29) and (30) are then used for determining σe . Having determined this quantity,
we calculate ge using Eq. (27).
To test the accuracy of these two approximations, let us consider a simple
case, namely, a resistor network with a percolation-type conductance distribu-
3.1. Strongly Nonlinear Composites 71
tion, f (g) = (1 − p)δ(g) + pδ(g − 1). In this limit, Eq. (25) reduces to (Sahimi,
1993a)
(pZ/2)n − 1 1/n
ge = , (31)
(Z/2)n − 1
while Eqs. (29) and (30) predict that (Sahimi, 1993a)
$ %
p − 2/Z 1/n
ge = p(n −1)/n
2
. (32)
1 − 2/Z
Equations (31) and (32) do meet the two criteria that we set above, namely, that they
both reduce to the linear EMA for n = 1, and their predictions for the percolation
threshold are the same as in the case of linear transport: Both equations predict
that ge vanishes at p = pc = 2/Z, the same as that predicted by Eq. (24) for linear
transport. We can also compare the predictions of these EMAs with those for the
effective microscopic conductivity of the Bethe lattice. For example, for n = 1/2
Eq. (7) predicts that µn = 3, whereas the numerical estimate for 3D systems (see
below) for n = 1/2 is µn 2.35. However, unlike the two EMAs described above,
the region near pc in which the conductivity of a Bethe lattice is different from
that of a 3D network is so narrow that it can hardly be detected (see Figure 3.2).
Consider now the case in which the nonlinear composite material obeys a
current-field response of the following form
I = g|E|1/n E, (33)
which is a slight generalization of Eq. (1). Bergman (1989) and Lee and Yu (1995)
developed an EMA for computing the effective conductivity of this type of com-
posite materials. Bergman developed an EMA for any value of n, while Lee and Yu
considered only the n = 1/2 limit. In both cases a 2D continuum model (but with
percolation disorder) in which inclusions, consisting of long cylinders (or circles)
of nonlinear conductance gα (α = i, h), representing the inclusion and the host
matrix, were embedded in an effective medium with a nonlinear conductance ge .
As usual, one applies a uniform far field E0 , calculates the local field Eα , and in-
sists that Eα = E0 . We supplement Eq. (33) by the usual electrostatic equations,
namely,
∇ · I = 0, ∇ × E = 0. (34)
Then, there exists a potential ϕ such that
E = −∇ϕ. (35)
If the potential ϕ is known, then, one can calculate Eα . Trial functions of the
following form,
ϕα (r, θ ) = −E0 (1 − bα )r cos θ, r < R, (36)
are now selected, where bα is a variational parameter, and R is the radius of the
cylinder. With these choices, the energy functional of the composite is given by
$ %
1
Hα = ge + pα ge −1 + 4bα + 4bα2 + bα4 + pα ge (1 − bα )4 V04 , (38)
3
where pα is the volume fraction of material of type α. If we now define yα = gα /ge
and minimize the energy functional, we obtain
(1 + yα )bα3 − 9yα bα2 + 3(2 + 3yα )bα + 3(1 − yα ) = 0, (39)
which provides an equation for bα and ϕα , and hence Eα . If the system is such
that inclusions of nonlinear conductivity gi and volume fraction pi are randomly
distributed in a host of conductivity gh with volume fraction ph (pi + ph = 1.0),
then the EMA equation is simply given by
pi bi (yi ) + ph bh (yh ) = 0. (40)
Figure 3.3 compares the predictions of this EMA with the results of numerical
simulation, demonstrating the accuracy of the predictions.
0.0
-0.5
-1.0
[ ]
(n)
ge -1.5
log
(l )
ge
-2.0
-2.5
-3.0
F
(n)
Figure 3.3. Effective nonlinear conductivity ge , normalized by the effective conductivity
of the system in the linear regime, versus the fraction p of the good conducting bonds.
Solid curves are the predictions of the EMA, Eqs. (39) and (40), while symbols show the
results of numerical simulations. The results are, from top to bottom, for conductivity ratios
y = 0.5, 0.1, 0.01 and 0.001 (after Lee and Yu, 1995).
3.1. Strongly Nonlinear Composites 73
so that the right-hand side of Eq. (41) is only a function of |E|2 i . Therefore,
Eq. (41), when written for both phases a and b, forms a set of coupled self-
consistent equations, the solution of which yields E 2 a /E02 and E 2 b /E02 . Given
these two quantities and Eq. (29), the effective generalized conductivity ge is then
estimated.
As an example, consider a 2D system. With f (σ ) = pδ(σ − σa ) + (1 − p)
δ(σ − σb ) and Z = 4, Eq. (29) yields
σe
ge = 1/n
E0
1/2
1 1/2n
= 1/n
(1 − 2p)(Xb − Xa ) + (1 − 2p)2 (Xb − Xa )2 + 4Xa Xb , Xi = gi |E|2 i ,
2E0
(43)
and from Eq. (41) one obtains, for example,
E02 2Xb − (1 − 2p)2 (Xb − Xa )
E a =
2
(2p − 1) + . (44)
2p [(1 − 2p)2 (Xb − Xa )2 + 4Xa Xb ]1/2
It can then be shown that Eq. (43) is identical with the Hashin–Shtrikman lower
bound for ge , derived by Ponte Castaneda et al. (1992) and described in Chap-
ter 2. Numerical simulations of the problem indicated close agreement with the
predictions of Eq. (43).
Two other methods that have been proposed for treating the problem of conduc-
tivity of a nonlinear material embedded in a matrix are the perturbation expansion
and the variational approach. Normally, these methods are described as part of the
74 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
continuum approach to these problems. However, since they were developed for
materials with percolation disorder, we describe them here, rather than in Chapter
2. What follows is a brief description of each method.
with respect to an arbitrary variation δϕ(x) away from the solution of Eq. (48),
provided that δϕ vanishes on the surface of the inclusions. When the minimum
condition is satisfied by a trial function ϕ̂, the effective nonlinear conductivity is
obtained from
ge E0 = Ĥ =
4
g(x)|Ê(x)|4 d, (50)
where Ê = ∇ ϕ̂. Thus, it remains to develop suitable trial potential functions ϕ̂.
The trial functions must be selected so as to satisfy the symmetry of the system
and the boundary conditions that are imposed on it. Thus, if the inclusions are
cylindrical, then, the trial functions, similar to Eqs. (36) and (37), are expansions
in cos mθ (with m = 1, 3, 5, · · ·), whereas for spherical inclusions one must use
Legendre functions. If the trial functions are selected to be Eqs. (36) and (37)
(which involve only the parameter bα ), then Eq. (39) is obtained again. Yu and Gu
(1995) improved the accuracy of the method by using higher-order terms in the
expansions. Hence, for a cylindrical inclusion of radius R, they used
ϕi (r, θ ) = (c11 r + c13 r 3 /R 2 + c15 r 5 /R 4 ) cos θ
varies from phase to phase, but the exponent n is the same for all the components.
The duality relations that we describe here is due to Levy and Kohn (1998).
Consider a two-phase composite in which the local conductivities are defined
by
gj (|V|) = gj |V|1/n , j = 1, 2. (53)
The dual composite is another two-phase material with the same morphology, but
with phases that have the following local conductivity,
1 −n/(n+1) −1/(n+1)
gj (|I|) = = gj |I| , j = 1, 2. (54)
gj (|V|)
The effective conductivities of the two materials are expressed as
However, ij < I , and therefore (ij /I )n should vanish as n → ∞, implying that the
blob resistance will be zero, and thus all of the resistance of the cluster (material) is
offered by the red bonds, hence proving Eq. (67). By similar arguments Blumenfeld
and Aharony (1985) also proved that
ζ̃ (n = 0+ ) = Dmin , (69)
3.1. Strongly Nonlinear Composites 79
where Dmin is the fractal dimension of the minimum or chemical path between
two points of a percolation cluster, i.e., the shortest path between the two points.
Thus, for L < ξp , the minimum length Lmin scales with L as, Lmin ∼ LDmin , with
Dmin 1.13 and 1.34 in 2D and 3D, respectively. Moreover, Blumenfeld et al.
(1986) showed that
ζ̃ (n = 0− ) = Dmax . (70)
Here Dmax is the fractal dimension associated with the longest self-avoiding walk
(that is, a random walk in which the walker never visits any point more than once)
between the two terminals of the percolation network; if Lmax is the length of the
walk, then Lmax ∼ LDmax . Blumenfeld et al. (1986) also proved that
ζ̃ (n = −1) = Dbb , (71)
with Dbb being the fractal dimension of the backbone of percolation clusters. Note,
however, that it has not been possible to relate ζ (n = 1) to any of the topological
exponents. Blumenfeld et al. (1986) also proved that ζ (n) decreases monotonically
with n, and therefore dζ (n)/dn ≤ 0, with the equality holding at n = ∞. Using
values of the various exponents and fractal dimensions given in Table 2.3 of Volume
I, we see that in 2D, ζ (n = ∞) = 1, and ζ (n = −1) 2.18, whereas in 3D ζ (n =
∞) = 1, and ζ (n = −1) 1.6. Therefore, ζ (n) is a slowly-varying function of n.
In addition to direct numerical simulations, there are at least two other methods
for estimating µn and its dependence on n. These methods are generalizations of
those discussed in Chapter 5 for the linear conductivity, and in what follows we
describe them briefly.
where sij = 1 if the two sites i and j belong to the same percolation cluster and
sij = 0 otherwise, and the averaging is over all configurations of the occupied
sites (probability p) and unoccupied ones [probability (1 − p)]. We now define a
resistive susceptibility χR (n; C) for a cluster C of sites via
χR (n; C) = Rij (n), (73)
i∈C j ∈C
where Rij (n) is the nonlinear resistance between sites i and j . Then, the total
resistive susceptibility χR (n), defined by
, -
χR (n) = Rij (n) , (74)
j
80 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
is obtained by summing χR (n; C) over all cluster, weighting each cluster by its
probability of occurrence. This is usually done in terms of cumulants, whereby
one writes
χR (n) = N (C; d)pnb (C ) χRc (n; C). (75)
C
In this equation nb (C) is the number of bonds in the cluster, N (C; d) is the number
of ways per site a diagram, topologically equivalent to C, can be realized on a d-
dimensional simple-cubic lattice, and the sum is over all topologically inequivalent
diagrams C. Moreover, χRc (n; C) is the cumulant defined by
χRc (n; C) = χR (n; C) − χRc (n; γ ), (76)
γ ∈C
where the sum is over all subdiagrams γ of C. Then, the average resistance R(n)
is defined by
χR
R(n) = ∼ |p − pc |−ζ (n) . (77)
χp
Therefore, the procedure for series analysis of resistance of random resistor
networks with power-law conductors is as follows. For each cluster C, and fixed
values of n (the power-law exponent) and nb (the number of bonds in the cluster),
the resistance Rij (n) is computed (by solving the Kirchhoff’s equations). These
computations are carried out for all such clusters, from which χR (n; C) and hence
χR (n) are obtained. Writing
χR (n) = A(k, l)d l p k , (78)
k l
one obtains a power series in p for χR (n). Since, in practice, the number of possible
cluster configurations increases very rapidly with nb , the computed power series
cannot be very long. For example, Meir et al. (1986) calculated the first 11 terms
of the series. Another series is obtained for χp , the computation of which is very
simple since it involves only counting of the number of clusters’ configurations.
The resulting two power series are then analyzed by a Padé approximation method,
from which the average resistance R(n) and hence the resistivity exponent ζ (n)
are computed. Using the results of Meir et al. (1986) and Eq. (66), we present
in Figure 3.4 the variations of µn = µ(n) with n. This figure indicates that µn
decreases very rapidly with increasing.
Figure 3.4. Dependence of the power-law conductivity exponent µ(n) on the power-law
exponent n (after Sahimi, 1993a, plotted based on the results of Meir et al., 1986).
Since (Harris et al., 1975), ν = 1/2 + 5/84 + O( 2 ), we obtain, using Eq. (66),
5 17 1 1 7(n − 1)
µ(n) = − + − 3+ , (80)
2 84 n 2 84 36
which reduces, in the limit n = 1, to Eq. (5.233) of Volume I for linear conductivity.
Such -expansions, while predicting the correct general trends in the n- and d-
dependence of the exponent µ(n), are not very accurate for the practical cases of
d = 2 and 3.
noise depends on the morphology of the conducting sample. Resistance noise was
studied in Chapter 5 of Volume I for the case of linear composites. In this section,
we consider the same problem for power-law conductors described by Eq. (1).
Consider a sample composite in which each of the conducting elements of the
nonlinear resistors has the same average value, but is fluctuating independently
with a correlation δra δrb = ρ 2 , where ra and rb are two resistances. Then, similar
to what was discussed in Chapter 5 of Volume I, the relative noise SR is calculated
from
2(n+1)
δRδR ρ2 b ib
SR = = , (81)
R2 r 2 ( b ibn+1 )2
where R is the resistance of the sample, ib is the current in the bonds, and
the sums are over all the current-carrying bonds. Note that the voltage noise
SV = δV δV /V 2 itself is given by, SV ∼ I 2n , and that for a homogeneous,
d-dimensional lattice of linear size L, SR = (ρ 2 /r 2 )/Ld .
For the sample-spanning percolation cluster at pc (or, equivalently, at length
scales L < ξp above pc ) the resistance noise scales with L as
SR ∼ L−bn , (82)
where bn = b(n) is the analogue of the exponent b for linear conduction,
Eq. (5.250) of Volume I. Rammal and Tremblay (1987) showed that
ζ̃n ≤ bn ≤ Dbb , bn ≤ 2ζ̃n − Dr , (83)
where, as before, ζ̃n = ζ̃ (n) = ζ (n)/ν, and Dbb and Dr are the fractal dimensions
of the backbone and the red bonds, respectively. As discussed in Chapter 5 of
Volume I, these bounds are also satisfied in the linear conduction case. Near the
percolation threshold pc ,
SR ∼ (p − pc )−κn , (84)
where, similar to the case of linear conduction, κn = κ(n) is a completely new
exponent independent of all the percolation exponents. Of course, κn and bn are re-
lated, κn = ν(d − bn ), and therefore the above bounds for b(n) can be immediately
converted to bounds for κn .
While SR is related to the 4th moment of the current distribution, one can, similar
to linear conduction discussed in Chapter 5 of Volume I, construct the general
moments Mq (x, x ) of the current distribution between two points x and x ;
(n+1)q
Mq (x, x ) = ib , (85)
b
where, as before, the sum is over all the current-carrying bonds of the network.
Then, for self-similar morphologies, such as the sample-spanning percolation clus-
ter at pc (or at length scales L < ξp above pc ), one can define an infinite hierarchy
of exponents τq (n) for |x − x | ∼ L:
Mq (x, x ) ∼ L−τ̃q (n) , (86)
3.2. Nonlinear Transport Caused by a Large External Field 83
Similar to the case of linear conduction, the exponents τq (n) are independent of
each other. Moreover, Rammal and Tremblay (1987) proved that τ0 − τq (n) is a
decreasing convex function of q that satisfies the following inequalities,
q 1
τq−1 (n) ≤ τq (n) ≤ τq−1 (n) − τ0 , (87)
q −1 q −1
where the last of the two inequalities is valid only for q ≥ 1. For the sample-
spanning percolation cluster at length scales L < ξp , one has, τ̃q (n) = τq (n)/ν.
Rammal and Tremblay (1987) obtained approximate (but not particularly accurate)
estimates of these exponents.
Figure 3.5. A strong external potential induces dynamic anisotropy in a material, giving
rise to two correlation lengths ξL and ξT . Circle denotes the point at which the potential is
applied to the system (after Sahimi, 1993a).
ξT ∼ |F − Fc |−νT . (89)
The problem studied here has certain similarities with directed percolation (Kinzel,
1983; Duarte, 1986,1990,1992; Duarte et al., 1992). In directed percolation, the
bonds of a network are directed and diode-like. Transport along such bonds is
allowed only in one direction. If the direction of the external potential is reversed,
then there can be no macroscopic transport in the new direction. Similar to the
present problem, in directed percolation one also needs two correlation lengths to
characterize the shape of the percolation clusters. However, there is an important
difference between what we study here and directed percolation: The anisotropy
in our system is dynamically induced, whereas the bias and anisotropy in directed
percolation are static and fixed.
An example of such nonlinear systems is the model proposed by Narayan and
Fisher (1994) (see also the somewhat related model proposed by Herrmann and
Sahimi, 1993, and Herrmann et al., 1993). They considered a randomly-rough
surface onto which a fluid or a charge carrier is poured into isolated “lakes,” such
that initially a sample-spanning cluster of connected lakes does not exist. The
surface is then slowly tilted at an angle θ , such that the fluid spills out of the filled
lakes and feeds unfilled lakes further downhill. For θ < θc , where θc is the critical
value of the tilt angle, the filled lakes cluster together. The characteristic size of
such clusters increases as θ does, and diverges at θ = θc . Above θc the system
becomes depinned, so that the fluid or the charge carrier can flow from the top
to the bottom of the system. Near and above θc the transport process is highly
inhomogeneous and confined to narrow and well-separated channels, somewhat
3.2. Nonlinear Transport Caused by a Large External Field 85
similar to Figure 3.5. Note that, under the influence of gravity, a force builds up
at the terminus of a cluster, rather than being uniform everywhere in it. Therefore,
when θ increases, clusters grow from their terminus sites, with a higher probability
of growing if they are already large. This implies that, the dominating flow paths
cannot be determined by a local analysis that searches for weak links in the system.
Rather, one must consider the entire system, i.e., the phenomenon is non-local.
The above description is a continuum one, but has a well-defined lattice coun-
terpart. In the lattice model, the sites represent the lakes, while the bonds are the
transport paths that connect the lakes. A force F is imposed on the lattice, and it
suffices for each site i to have outlets connecting it only to its d nearest neighbors
iα in the next plane downhill, where d is the dimensionality of the system. It is
assumed that the current flowing in a path depends only on the depth above the
lip of the lake it emerges from. Thus, a barrier biα is assigned to each outlet α
emerging from a site i which controls the current flowing through the outlet. The
barriers are selected randonmly and independently from a distribution. At each
site i of the lattice there is a depth of fluid hi . The current Iiα flowing through an
outlet α from a site i is zero if hi < biα − F , and
Iiα = (hi − biα + F ) if hi > biα − F. (90)
The exponent characterizes the transport over the barrier lip. Narayan and Fisher
(1994) presented arguments that indicate that = 3 + d/2 for a d-dimensional
system. Note that an increase in F is equivalent to uniformly lowering all the
barriers biα . √
Narayan and Fisher (1994) argued that ξT ∼ ξL . That is, we can imagine that
the consecutive events in which the bonds are filled with the flowing current are
in fact consecutive steps of a random walk in the (d − 1) transverse directions.
If so, the longitudinal direction acts as the time axis, and therefore the distance
that the random walker travels in the transverse direction should increase with √ the
square root of time (the usual law of random walks), implying that ξT ∼ ξL ,
and thus νT = νL /2. The random-walk argument can also be used to estimate
the upper critical dimension du of the system at and above which the mean-field
theory is exact. The clusters perform random walks in the (d − 1)-dimensional
transverse space, with the longitudinal direction acting as the time coordinate.
From the theory of random walks (Hughes, 1995) we know that if d − 1 > 2, then
two walks that start out close to each other have a finite probability of not crossing
each other, whereas for d − 1 < 2 they are certain to cross. Therefore du − 1 = 2
and hence du = 3. This immediately implies another significant difference between
this model and directed percolation for which du = 5 (Obukhov, 1980), and also
with isotropic percolation for which du = 6.
Narayan and Fisher (1994) studied various topological and transport properties
of this model. One surprising aspect of this phenomenon is that the critical expo-
nents that characterize the power-law behavior of the properties of interest above
and below, but near, the threshold Fc are not equal. Consider first the system be-
low the threshold. We write ξL ∼ |F − Fc |−νL , where superscript b signifies the
b
fact that the critical exponent is associated with the regime below the threshold. The
86 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
() (n)
where ge and ge are the effective linear and nonlinear response of the network,
respectively. In general, as our discussion throughout this book should have made
()
it clear, the effective linear conductivity ge in a binary random network with
components gA and gB can always be written as
ge() = F (gA , gB , p), (108)
where F is a function which, in general, depends on the geometry of the system.
(n)
Then the effective nonlinear response ge of the system is given by
2
g (n) ∂F
ge =
(n)
()
. (109)
p ∂ge
That is, the effective nonlinear response is estimated based on an estimate of the
effective conductivity of the same material but in the linear regime. Recall from
Chapter 2 that the same sort of idea was developed by Ponte Castañeda (1992b) in
the context of the continuum models. The derivation of Eq. (109) will be discussed
in detail in Section 3.4, where we describe the derivation of a similar equation for
the dielectric constant of the same type of composites. Therefore, if the function
(n)
F can somehow be calculated, ge will also be determined from Eq. (109). Since
F is an estimate of the effective conductivity of a linear binary composite, we
may use the EMA, Eq. (24) (or, for example, the Maxwell–Garnett or any other
approximation), for linear resistor networks which for our binary network with
f (g) = pδ(g − gB ) + (1 − p)δ(g − gA ) is given by
() ()
gA − g e g B − ge
(1 − p) ()
+p ()
= 0. (110)
gA + ge (Z/2 − 1) gB + ge (Z/2 − 1)
(n)
Thus, the procedure for calculating ge by an EMA is as follows. One first solves
() ()
Eq. (110) for ge . This equation, which is quadratic in ge , defines the function F .
() (n)
Having determined ge , one utilizes Eq. (109) to calculate ge . Figures 3.6 and 3.7
compare the results of computer simulations in the square network in two limiting
cases with the EMA predictions. The numerical results in Figure 3.6, which are
for gA = 10, gB = 20, and g (n) = 0.1, are in excellent agreement with the EMA
predictions. The reason for the agreement is that the difference gB − gA is not large
and thus, as discussed in Chapter 5 of Volume I, the function F (i.e., the EMA
()
estimate) provides accurate predictions for ge . On the other hand, the numerical
results shown in Figure 3.7, which are for gA = 5000, gB = 10, and g (n) = 0.1,
agree only qualitatively with the EMA predictions because, as discussed in Chapter
5 of Volume I, in this case, due to the large difference between gA and gB , F (i.e.,
()
the EMA estimate) cannot provide accurate predictions for ge , which is consistent
with the general properties of the EMA.
It is clear that the development of an EMA for this class of composites involves
two stages. More generally, one may consider composites with more complex
I − V characteristics and develop a similar, but multistage, procedure for an EMA-
based computation of their effective transport properties. For example, one may
90 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
1.0
0.8
0.6
(n)
ge
(l )
g e 0.4
0.2
0.0
0.0 0.2 0.4 0.6 0.8 1.0
F
(n)
Figure 3.6. The effective nonlinear conductivity ge , normalized by the effective conduc-
tivity of the system in the linear regime, versus the fraction p of nonlinear conductors in
the square network. Solid curve shows the EMA predictions, while the symbols show the
results of numerical simulations for gB = 2gA = 20 and g (n) = 0.1 (after Yang and Hui,
1991).
250
200
150
(n)
ge
(l )
ge
100
50
0
0.0 0.2 0.4 0.6 0.8 1.0
F
Figure 3.7. Same as in Figure 3.6, but for gA = 5000, gB = 10, and g (n) = 0.1 (after Yang
and Hui, 1991).
3.3. Weakly Nonlinear Composites 91
consider a composite material (Yu and Gu, 1993) in which a fraction p of the
system has an I − V characteristic given by, i = gB v + gn1 v 3 + gn2 v 5 , while the
rest of the composite, with fraction 1 − p, is made of linear conductors, i = gA v.
One may compute the effective linear and nonlinear response of such a composite
() (n1) (n2)
defined by, I = ge v + ge v 3 + ge v 5 , by first solving the EMA equation for
()
the effective linear conductivity of the composite ge . Then, an equation similar
(n1)
to (108) is used for computing the first nonlinear conductivity ge . The two
() (n1)
conductivities ge and ge so obtained are then used in a higher-order equation
(n2)
in order to compute ge .
where ib is the current in bond b, which depends implicitly on n, and the sums
are over all the conducting bonds of the network. Blumenfeld et al. (1986) had
already proved that
.
∂P .. 1 0 n+1
. = |ib | , (112)
∂n rn =0 n+1
b
where ib0 = ib (rn = 0). Therefore, to linear order in rn , we can replace ib by ib0
and write
1 rn
P = r M1 I 2 − M(n+1)/2 I n+1 , (113)
2 n+1
where I is the total current in the network, and
2q
i0
Mq = b
, (114)
I
b
is the 2qth moment of the current distribution in the linear random resistor net-
work. As already discussed in Section 3.1.8 [see Eq. (86)] for the case of strongly
nonlinear composites, and in Section 5.16 of Volume I for linear systems, for
L ξp the moments of the current distribution scale with L as
Mq ∼ L−τ̃q , (115)
92 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
where all the τ̃q s are distinct. This means that the current distribution in a linear
random resistor network is multifractal, i.e., each of its moments scales with L
with a distinct exponent, which is similar to the moments of the force distribution
in elastic and superelastic percolation networks described in Chapter 8 of Volume
I (see Stanley and Meakin, 1988, for a review of general properties multifractal
()
systems and distributions). Therefore, the effective linear resistance Re of the
()
network, which is obtained via Re = ∂ 2 P/∂I 2 , shows deviations from a constant
value for n > 1 and
1/(n−1)
M1
I > Ic (L) ∼ ic ∼ ic Ly τ̃1 ∼ [ge() (L)]−y , (116)
M(n+1)/2
and therefore (Aharony, 1987)
τ̃(n+1)/2
1−
τ̃1
y(n) = . (117)
n−1
Since τ̃q is a monotonic and convex function (see, for example, Blumenfeld et
al., 1986), so also is y(n). For example, for d = 2 and 3 one has y(3) 0.08 and
0.06, and y(0) 0.18 and 0.1, respectively. This means that 0 < y(n) < y(1), and
therefore the linear regime I < Ic (L) extends to larger currents for larger linear
sizes L, implying that even a narrow nonlinear regime will be enhanced (see also
below) in a percolation network. A similar analysis for L ξp yields (Aharony,
1987)
d − 1 − y(n)τ̃1
x(n) = , (118)
d − 2 + τ̃1
and therefore for d = 2 one finds that x(n) = 1.03 − y(n). Since y(n) > 0,
Eq. (118) does not agree with the experimental result of Gefen et al. (1986) for
any n, and therefore a simple percolation network in which each conducting bond
follows Eq. (104) cannot explain Gefen et al.’s data.
To study scaling properties of weakly nonlinear composites near the percolation
threshold pc , we must consider resistance and conductance fluctuations in linear
resistors networks. Recall from Section 5.16 of Volume I that, for a percolation net-
()
work near pc , the relative linear resistance noise, SR = δRδR/[Re ]2 , follows
the following power law [see also Eq. (84) for strongly nonlinear composites]
SR ∼ (p − pc )−κ , (119)
which defines the critical exponent κ. One can, in a similar fashion, consider con-
ductance fluctuations SG of a linear superconducting percolation network below
pc . In this case
SG ∼ (pc − p)−κ . (120)
It can be shown (Wright et al., 1986) that in 2D, κ = κ . Given Eqs. (119) and (120),
we can discuss some of the scaling properties of weakly nonlinear composites near
pc .
3.3. Weakly Nonlinear Composites 93
Stroud and Hui (1988) considered a composite with the following characteristic,
I(x) = g () (x)E(x) + g (n) (x)|E(x)|n E(x), (121)
where n ≥ 1, and g () and g (n) are the linear and nonlinear conductivities of the
medium, respectively, which depend, in general, on the spatial position x, and
the applied electric field (or voltage) E. Equation (121) is just another version
of (104), written explicitly for the current. As mentioned earlier, if one assumes
that all the components in the disordered composite have inversion symmetry, then
n = 2, which was the case studied by Stroud and Hui (1988). The volume-averaged
current I is defined by
I = ge() E0 + ge(n) |E0 |2 E0 , (122)
with E = E0 . Consider now the dissipated power for this composite which, in a
continuum formulation, is given (for n = 2) by
P = I · E d = ge() |E0 |2 + ge(n) |E0 |4 . (123)
This equation, in which is the volume of the composite, is the continuum analog
of Eq. (111). Using Eq. (121), we rewrite Eq. (123) as
P= g () (x)E · E + g (n) (x)|E|4 d = P2 + P4 . (124)
Then, to first order in g (n) (x), the second term of Eq. (124) is rewritten as,
P4 = g (n) (x)|E|4 = P4 , (125)
where the subscript indicates that the electric field must be calculated from
the solution of the linear problem, i.e., in the limit, g (n) (x) = 0. In reality, the
difference E − E is of first order in g (n) , and therefore will contribute to P4 only
a second-order term. By a similar argument, one can show that
P2 = P2 . (126)
() (n)
Therefore, to first order in g (n) (x), the effective conductivities ge and ge are
given by (Stroud and Hui, 1988)
1 g () |E |2
ge =
()
g ()
(x)|E | 2
d = , (127)
|E0 |2 |E0 |2
1 g (n) |E |4
ge(n) = g (n)
(x)|E | 4
d = . (128)
|E0 |4 |E0 |4
Observe that Eq. (128) is the same as (50) for strongly nonlinear composites. Equa-
tions (127) and (128) are manifestations of an important result: The effective linear
and nonlinear conductivities of a weakly nonlinear composite can be calculated
from the behavior of the electric field in the linear problem.
Utilizing a similar line of analysis, Stroud and Hui (1988) proved another im-
portant property of weakly nonlinear composites, namely, that to first order in
94 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
(n)
g (n) (x), ge is essentially given by the mean square conductivity fluctuations in a
linear composite,
[g () ]2
ge(n) = , (129)
c
where g () is the root mean square conductivity fluctuations in the linear compos-
ite, and c is a constant with dimensions of energy. Note that, since the conductivity
fluctuations cause corresponding fluctuations in the current, which in turn are
related to the 4th moment of the current distribution (see above), Eq. (129) is
consistent with, but much more general than, Aharony’s result, Eqs. (113)–(118),
discussed above.
Using Eq. (129), one can now deduce the power-law behavior of the nonlinear
(n) (n)
conductivity ge near the percolation threshold pc . According to Eq. (129), ge is
given by conductivity, or resistivity, fluctuations of the linear conductivity problem.
Therefore (Stroud and Hui, 1988), using Eq. (119), we can write
(n)
ge
()
∼ (p − pc )−κ , (130)
[ge ]2
which, when combined with the power-law behavior of the effective linear
()
conductivity ge near pc , Eq. (18), yields
ge(n) ∼ (p − pc )2µ−κ , (131)
where µ is the critical exponent of the effective linear conductivity near pc . Note
that in a composite in which a fraction p of the material is superconducting and
the rest is made of weakly nonlinear conducting material, one has
ge(n) ∼ (pc − p)−2s−κ . (132)
With the help of Eqs. (131) and (132), one can construct a general scaling repre-
sentation for the effective conductivity of a composite, a fraction pM of which is a
() (n)
good weakly nonlinear conductor characterized by, I = gM V + gM V 3 , while the
rest of the composite, with a fraction (1 − pM ), is a poor weakly nonlinear conduc-
() (n) () () (n) (n)
tor which follows, I = gI V + gI V 3 , with gM gI and gM gI . Then,
() (n)
with z = [gI /gM ]/(p − pc )µ+s , p = |p − pc |, and considering Eqs. (131)
and (132), one can write (Levy and Bergman, 1994b)
ge(n) gI p −2s−κ I (z) + gM p 2µ−κ M (z).
(n) (n)
(133)
The properties of the two scaling functions I and M vary in three distinct
regimes.
(1) In regime I, which is for pM > pc and |z| 1, the scaling function M must
be constant in order for one to be able to obtain Eqs. (130) and (131). It is then
straightforward to see that I must also be a constant.
(2) In regime II, which is for pM < pc and |z| 1, the scaling function I
must be constant so that one can recover Eq. (132). It is not difficult to see
3.3. Weakly Nonlinear Composites 95
We may interpret Eq. (136) as meaning that, the exponent x defined by Eq. (102),
is given by, x = 12 (1 + κ/µ). In 2D, where κ 1.12, Eq. (136) predicts that,
x 0.93, which still does not agree with Gefen et al.’s measurement, x 1.47,
but is closer to it than the prediction of Eq. (118).
96 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
The foregoing scaling laws are valid when one has cubic nonlinearity, i.e., when
n = 2 in Eq. (121). Zhang (1996a) and Gao et al. (1999) generalized these results
to any n. For the first limiting case, i.e., a composite of insulating and weakly
nonlinear conducting materials near pc , Gao et al. (1999) obtained the following
estimates,
νd − ζ − µ 1 + (νDbb − 1)−n/2 (ζ − 1)n/2+1 − ζ
v= + , (146)
2 n
νd − ζ + µ 1 + (νDbb − 1)−n/2 (ζ − 1)n/2+1 − ζ
w= + , (147)
2 n
where Dbb is the fractal dimension of the backbone of the percolation cluster,
and ζ is the resistivity exponent defined by Eq. (65). For the case of a composite
of superconducting and weakly nonlinear conducting materials, Zhang (1996a)
obtained the following estimate,
s 12−n κ [(n + 2)/2]
v = + νd + , (148)
2 2 n n
where κ [(n + 2)/2] is the exponent associated with the conductance fluctuations
below the percolation threshold defined above. In the limit n = 2 Eq. (148) reduces
to (142). Numerical simulations for testing the validity of these predictions were
reported by Levy and Bergman (1993, 1994b) and Zhang (1996b).
the electric field E). The local conductivities of the two-phase material are given
by
() (n)
gj (V) = gj + gj |V|n , j = 1, 2. (149)
The dual composite is another two-phase material with the same microgeometry,
but with its phases having the following local conductivity,
(n)
1 1 gj
gj (|I|) = = () − () |I|n . (150)
gj (|V|) gj [gj ]n+2
The effective conductivities of the two components can be expressed as
being the magnitude of the current that flows through the primal composite, which
is also the magnitude of volume-averaged electric field in the dual composite.
All the notations have the same meaning as for the strongly nonlinear composites
discussed earlier. To first order in the local nonlinear conductivity g (n) , the effective
conductivities satisfy
(n,d)
1 ge
ge() + ge(n) V0n = (,d)
− I n.
(,d) 2 0
(154)
ge [ge ]
This relation leads us to
1
ge() = (,d)
, (155)
ge
which is the same as the well-known duality relation for linear composites, and
(n,d)
ge
ge(n) V0n = − I n.
(,d) 2 0
(156)
[ge ]
Equation (156) implies immediately that for cubic nonlinearity (n = 2),
(n) (n,d)
ge ge
()
=− (,d) 2
. (157)
[ge ]2 [ge ]
Similar to the case of strongly nonlinear composites described in Section 3.1.6,
we can extend this analysis to weakly nonlinear materials near the percolation
threshold and investigate its consequences. As discussed in Section 3.3.2, if we
3.3. Weakly Nonlinear Composites 99
() (n)
have a mixture of good conductors [conductances gM and gM ] and perfect
insulators, then near pc we expect to have [see Eq. (130)]
(n) (n)
ge gM
()
∼ ()
(p − pc )−κ , (158)
[ge ]2 [gM ]2
where κ is the exponent for the resistance noise introduced and described above.
() (n)
Similarly, for a mixture of normal conductors [conductances gI and gI ] and
superconductors near pc , one must have [see Eq. (132)]
(n) (n)
ge gI
()
∼ ()
(pc − p)−κ , (159)
[ge ]2 [gI ]2
where the exponent κ was also defined above. Using the duality relations described
above, one can then show that κ = κ which, as discussed above and in Chapter 5
of Volume I, also holds for linearly conducting composites.
The foregoing discussions can be extended to the case in which the ratio gI /gM ,
for both the linear and nonlinear conductivities, is finite. In this case Eq. (133)
should hold for the primal composite and its dual, both above and below the
percolation threshold pc . Then, using the above duality relations, one can show
that the scaling functions ±,I and ±,M [where the plus (minus) sign is for
p > pc (p < pc )] and their dual counterparts satisfy the following relations
(d) (d)
M = I , I = M , (160)
with the understanding that if the left-hand side of Eqs. (160) uses the scaling
function with the plus sign, then, the right-hand side uses the function with the
minus sign, and vice versa.
Figure 3.8. A typical nonlinear I − V curve for a PrBa2 Cu3 O7−δ compound at 300 K.
Symbols show the data for four different samples (after Lin, 1992).
Figure 3.9. Logarithmic plot of the critical current Ic versus the effective linear conductivity
(l)
ge for four samples of PrBa2 Cu3 O7−δ . Solid circles show the data for an Ag-] added
(l)
sample. The straight line represents Ic ∼ [ge ]0.6 (after Lin, 1992).
3.4. Dielectric Constant of Weakly Nonlinear Composites 101
µ 2.5 and κ 5.14 for the 3D Swiss-cheese model, i.e., the model in which
spherical inclusions are distributed randomly in a uniform matrix, then Eq. (133)
predicts that x 0.74, which is only about 20% larger than Lin’s measurements
which, given the scatter in the data shown in Figure 3.9, is quite acceptable.
which is the analogue of Eq. (108). Here p1 is the volume fraction of the 1
component, and F is an estimate of the effective dielectric constant which, in
general, depends on the morphology of the composite. We initially assume that
()
only component 1 is nonlinear, so that 2 = 2 , and therefore we can invoke an
approximate nonlinear form of Eq. (163):
e = F (1 , 2 , p1 ), (164)
() (n)
where, i = i + i |Ei |2 , and |Ei |2 is the mean square of the electric field
in the ith component in the linear limit. We must keep in mind that Eq. (164) is
valid only if 1 and 2 are constant in their respective component, implying that E
is uniform in the nonlinear component.
102 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
()
The function F is now expanded in a Taylor series around e :
e F 1 , 2 , p1 + F 1 , 2 , p1 1 |E1 |2 ,
() () () (n)
(165)
where F = ∂F /∂1 . However, one can express F exactly in terms of the average
squared electric field in component 1 in the linear limit:
|E1 |2
()
∂e
() ()
p1 = ≡ F 1 , 2 , p1 , (166)
E02 ∂1
where E0 is the external field. Therefore,
(n)
1
e = e() + F |F |2 E02 , (167)
p1
which means that, by the definition of the effective nonlinear dielectric function
(n)
e , we obtain
(n) $ % . .
1 ∂e .. ∂e ..
e =
(n)
. (168)
p1 ∂1 . ∂1 .
Equation (168) is the analogue of Eq. (109) for the nonlinear conductivity.
We can generalize this result to composites in which both components are weakly
nonlinear. Hence, we write
(n) (n)
1
e = e() + F |F |E 2 + 2 F2 |F2 |E02 , (169)
p1 1 1 0 p2
where Fi = ∂e /∂i (i = 1, 2). Therefore,
(n) (n)
e(n) = 1 F1 |F1 | + 2 F2 |F2 |. (170)
p1 p2
Equation (170) also suggests an analogous generalization for the effective nonlin-
(n)
ear conductivity ge , which would then represent a generalization of Eq. (109).
One can now use this general method of approximation and study its properties in
certain limits.
The second morphology for which the effective nonlinear dielectric function can
be exactly computed is one in which the components are arranged in the form of
flat slabs perpendicular to the external field. For this case,
1
e() = () ()
, (173)
p1 /1 + p2 /2
from which one obtains, using Eq. (170),
(n) (n)
1 2
e(n) = p1 () ()
+ p2 () ()
. (174)
[p1 + 1 p2 /2 ]4 [p2 + 2 p1 /1 ]4
A perturbation expansion, similar to what we described in Section 3.1.4 for the
effective conductivity of strongly nonlinear composites, was also developed by Yu
et al. (1993).
Using Eq. (176), the functions Fi = ∂e /∂i are computed which, when sub-
()
stituted in Eq. (170), yield the MG estimate for the nonlinear dielectric function
104 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
(n)
e . Hui (1990a) extended the MG approximation to a more general composite
for which, D = () E + (n) |E|n E
We should emphasize again that in the type of nonlinear problems that we are
discussing here, the geometry of the system and the boundary conditions are very
important and have a profound influence on the overall behavior of the system.
As a matter of fact, every result described so far is valid only for two-terminal
systems, and essentially nothing is known for multi-terminal ones.
(1) If the applied electric field or current exceeds a critical threshold, then, at
zero frequency, strong nonlinearity results in the breakdown of the conducting
elements of a composite. The critical field decreases to zero as the volume
fraction of the conducting component approaches pc , hence indicating that
such composites become progressively more responsive to the external field
as pc is approached. This phenomenon is what we have referred to as threshold
nonlinearity; it will be studied in Chapter 5.
(2) Alternatively, although increasing the external voltage or current may not result
in electric or dielectric breakdown of a composite, it can lead to very large
enhancements of the nonlinearities as the volume fraction of the conducting
component approaches pc . We already described this phenomenon in Sections
3.1 and 3.3 in terms of the crossover from a linearly conductive material to a
weakly nonlinear one, and our goal in this section is to do the same for optical
properties of the same type of composite solids.
Following our discussions in Section 3.3, we consider in this section weak non-
linearities so that the field-dependent conductivity g(E) can be written as a power
series in the applied electric field E, with the leading term, i.e., the linear con-
ductivity g () , being much larger than the higher-order terms, a situation which is
typical of various nonlinearities in the optical and infrared spectral ranges of in-
terest to us. As discussed in Section 3.3, despite this weakness, such nonlinearities
lead to qualitatively new phenomena, such as enhancement of higher harmonics in
percolation composites, and the occurrence of bistable behavior of the composite
(Bergman et al., 1994; Levy et al., 1995) in which the conductivity switches be-
tween two stable values. In such disordered materials, especially those that contain
metal particles that are characterized by a dielectric constant with negative real
and small imaginary parts, the fluctuations in the local field are strongly enhanced
in the optical and infrared spectral ranges, leading to enhancement of various
nonlinear properties. If the disorder in the morphology of such solid materials is
of percolation-type, then they are potentially of great practical importance (see,
for example, Flytzanis, 1992) as composites with intensity-dependent dielectric
functions and, in particular, as nonlinear filters and optical bistable elements. The
optical response of such nonlinear composites can be easily tuned by, for example,
controlling the volume fraction and morphology of their constitutes.
More generally, optical properties of fractal aggregates of metal particles have
been studied. These studies indicate that a fractal morphology results in very large
enhancement of various nonlinear responses of the aggregates within the spectral
range of their plasmon resonances. The typical size, a ∼ 10 nm, of the metal
particles in such fractal aggregates is much smaller than the wavelength λ > 300
nm in the optical and infrared spectral ranges. Since the average density of particles
in fractal aggregates is much smaller than in non-fractal materials, and approaches
zero with increasing size of the aggregates, it is possible to consider each particle in
the aggregate as an elementary dipole and introduce the corresponding interaction
operator. If this is done, then, solving the problem of the optical response of
metal fractal aggregates reduces to diagonalizing the interaction operator for the
106 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
light-induced dipoles. If the size of the fractal aggregate is not very large, the
diagonalization can be done numerically (and efficiently) and thus the local electric
field can be calculated (see, for example, Stockman et al., 1995, 1996; Stockman,
1997; Shalaev et al., 1993; Markel et al., 1999). Computations of this type indicate
that large field fluctuations are localized in some small parts of the fractal aggregate
and change with the wavelength. These predictions and numerical computations
of large enhancements of optical nonlinearities in metal fractals have also been
verified experimentally for degenerate four-wave mixing and nonlinear refraction
and absorption. In these experiments, aggregation of silver particles (which were
initially isolated) into fractal clusters led to six orders of magnitude enhancement
of the efficiency of the nonlinear four-wave process and about three orders of
magnitude enhancement in the nonlinear refraction and absorption. The localized
and strongly fluctuating local fields in these fractal aggregates were imaged by
means of the near-field scanning optical microscopy (Shalaev et al., 1993; Markel
et al., 1999). A similar pattern was obtained for the field distribution in self-affine
thin films (Shalaev et al., 1996a,b; Safonov et al., 1998). As discussed in Chapter
1, such self-affine films possess a fractal surface with different scaling properties
in the plane of the film and normal to it.
Despite such progress, the distribution of the local field and the corresponding
nonlinearities were, until recently, poorly understood for metal-dielectric com-
posites with percolation-type disorder, especially in the most interesting spectral
range where the plasmon resonances occur in the metal grains. As shown in Sec-
tion 3.3, if a small volume fraction p 1 of a nonlinear material is embedded in
a linear host, the effective nonlinear response of the composite can be calculated
explicitly. As one may expect, the nonlinearities are enhanced at the frequency
ωr corresponding to the plasmon resonance of a single metal grain. Numerical
calculations (Stroud and Zhang, 1994; Zhang and Stroud, 1994) for a finite p also
indicate considerable enhancement in the narrow frequency range around ωr and,
moreover, the system sizes that can currently be used in the computations are not
large enough for drawing quantitative conclusions about the nonlinear properties
for frequencies ω ωr . However, we should recognize that a small system size
L may act as an artificial damping factor that cuts off all the fluctuations in the
local field when the spatial separation is larger than L, hence resulting in a cor-
responding decrease of the nonlinearities which may otherwise not be seen in a
large enough sample.
An alternative method to numerical simulations is the effective-medium approx-
imation (EMA) that has the virtue of mathematical and conceptual simplicity. We
already described in Sections 3.1 and 3.3 such EMAs for nonlinear composites
near pc . As discussed there, for the static case the predictions of the nonlinear EMA
(Wan et al., 1996; Hui et al., 1997) are in good agreement with numerical simu-
lations for 2D percolation composite. However, despite this success, application
of any type of nonlinear EMA is suspect for the frequency range corresponding
to the plasmon resonances in metal grains. This is due to the fact that both com-
puter simulations and experimental data for the field distribution in percolation
composites indicate that the distribution contains sharp peaks that are separated
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 107
by distances that are much larger than the metal grain size. Thus, the local electric
field cannot be assumed to be the same in all the metal grains of the composite,
implying that the main pillar of the EMA, i.e., the assumption of a uniform field,
fails for the frequency range corresponding to the plasmon resonance in the films.
To address this problem, a new theory of the distribution of the electromagnetic
field and nonlinear optical processes in metal-dielectric composites was devel-
oped (Sarychev and Shalaev, 1999; Sarychev et al., 1999). The theory is based
on the concepts of percolation processes, and takes advantage of the fact that the
problem of optical excitations in percolation composites can be mapped onto the
Anderson localization problem. It predicts localization of surface plasmons (SP)
in composites with percolation disorder, and describes in detail the localization
pattern. It also indicates that the SP eigenstates are localized on length scales that
are much smaller than the wavelength of an incident light. The eigenstates with
eigenvalues that are close to zero (resonant modes) are excited most efficiently
by the external field. Since the eigenstates are localized and only a small portion
of them is excited by the incident beam, overlapping of the eigenstates can typi-
cally be neglected, a fact that significantly simplifies the theoretical analysis and
allows one to derive relatively simple expressions for enhancement of linear and
nonlinear optical responses.
The purpose of this section is to describe and summarize this progress. An excel-
lent comprehensive review of this subject was presented by Sarychev and Shalaev
(2000). This section is patterned closely after their review and represents a sum-
mary of their discussions. Since the languages of nonlinear currents/conductivities
and nonlinear polarizations/susceptibilities, or dielectric constants, are completely
equivalent, they will be used interchangeably in this section.
Since Eqs. (179) and (180) are difficult to solve analytically, one discretizes them in
order to solve them by numerical simulations. If, for example, a standard 5-point (in
2D) or 7-point (in 3D) finite-difference discretization is used, then, a discrete model
on a simple-cubic lattice is obtained in which the metal and dielectric particles
are represented by metal and dielectric bonds of the lattice. Thus, Eq. (180), in
discretized form, takes on the form of Kirchhoff’s equations defined on a lattice.
Assuming that the external electric field E0 is directed along the z-axis, one obtains
ij (φj − φi ) = ij Eij (181)
j j
where φi is the electric potential at site i of the lattice, and the sum is over the nearest
neighbors j of the site i. For the bonds ij in the ±z-direction, the electromotive
force Eij is given by, Eij = ±E0 a0 (where a0 is the spatial period of the lattice),
while Eij = 0 for the other bonds that are connected to site i. Thus, the composite
material is modeled by a resistor-capacitor-inductor network in which the bond
permittivities ij are statistically independent and a0 is equal to the metal grain size,
a0 = a. In the case of a two-component metal-dielectric random composite, the
permittivities ij take values m and d with probabilities p and 1 − p, respectively.
To make further progress, we use a simple-cubic lattice which has a very large but
finite number of sites N and rewrite Eq. (181) in a matrix form:
Hφ = E, (182)
where φ = {φ1 , φ2 , . . . , φN }, and the elements of the vector E are, Ei = j ij Eij .
Here H is a N × N matrix such that for i = j , Hij = −ij = d > 0 and m =
(−1 + iκ)|m | with probabilities p and 1 − p, respectively, and H =
ii j ij ,
where . . j refers to nearest neighbors of site i, and κ is the usual loss factor, κ=
/ . . 1. The diagonal elements H are distributed between 2d and 2d ,
m m ii m d
where d is the dimensionality of the space.
Similar to the dielectric constant, we write H = H + iκH , where iκH rep-
resents losses in the system. The Hamiltonian H formally coincides with the
Hamiltonian of the problem of metal-insulator transition (Anderson transition) in
quantum systems, i.e., it maps the quantum-mechanical Hamiltonian for the Ander-
son transition problem with both on- and off-diagonal correlated disorder onto the
present problem. Hereafter, we refer to H as the Kirchhoff’s Hamiltonian (KH).
Thus, the problem of determining the solution of Kirchhoff’s equation, Eq. (181)
or (182), is equivalent to the eigenfunction problem for the KH, H n = n n ,
whereas the losses can be treated as perturbations.
Since m < 0, and the permittivity of the dielectric matrix is positive, the set
d
of the KH eigenvalues n contains eigenvalues with real parts that are equal (or
close) to zero. Then, eigenstates n that correspond to eigenvalues n such that,
|n | |m |, and |d |, are strongly excited by the external field and are seen as
giant field fluctuations, representing the resonant surface plasmon modes. If one
assumes that the eigenstates excited by the external field are localized, then they
should look like the peaks of the local field with the average distance between
110 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
them being about a(N/n)1/d , where n is the number of the KH eigenstates excited
by the external field.
Consider now the special case when m = − , which corresponds to the plas-
d
mon resonance of individual particles in a 2D system. Since a solution to Eq. (181)
does not change if m and d are multiplied by the same factor, we normalize the
system and set d = −m = 1. We also suppose, for simplicity, that the metal con-
centration is p = 0.5. In this case, the eigenstates n are all localized. On the other
hand, computer simulations (Müller et al., 1997) showed that there is a transition
from chaotic (Berry, 1977) to localized eigenstates for the 2D Anderson problem,
with a crossover region between the two. Consider first the case when the metal
volume fraction p = pc = 1/2 for the 2D bond percolation problem. Then, the
diagonal disorder in the KH is characterized by, Hii = 0, and H2 = 4, which
ii
correspond to the chaos-localization transition (Müller et al., 1997). Moreover, H
also possesses strong off-diagonal disorder, Hij = 0, which favors localization
(see, for example, Verges, 1998). There is therefore strong evidence that the eigen-
states n are localized for all n in the 2D system, although one cannot rule out
the possibility of inhomogeneous localization, similar to that obtained for frac-
tal clusters (see, for example, Stockman et al., 1994), or power-law localization
(Kaveh and Mott, 1981; Kramer and MacKinnon, 1993).
In the case of d = −m = 1 and p = 1/2, all parameters in H are of the
order of unity, and therefore its properties do not change under the transforma-
tion d ⇐⇒ m . Therefore, the real eigenvalues n are distributed symmetrically
around zero in an interval of the order of one. The eigenstates with eigenvalues
n 0 are effectively excited by the external field and represent the giant local
field fluctuations. When p decreases (increases), the eigenstates with eigenvalues
n 0 are shifted from the center of the distribution toward its lower (upper)
edge, which typically favors localization. Because of this, one may assume that in
2D the eigenstates, or at least those with eigenvalues n 0, are localized for all
metal volume fractions p.
The situation in 3D is much more complex. Despite the great effort and the
progress that has been made, the Anderson transition in 3D is not yet fully
understood. Computer simulations (Kawarabayashi et al., 1998) of Anderson lo-
calization in 3D [with d = −m = 1, p = 1/2, the diagonal matrix elements wii
distributed uniformly around 0, −w0 /2 ≤ wii ≤ w0 /2, and the off-diagonal ele-
ments wij = exp(iφij ), with phases φij also distributed uniformly in 0 ≤ φij ≤
2π ] show that in the center of the band the states are localized for the disorder
w0 > wc = 18.8. In the 3D H Hamiltonian discussed here, the diagonal elements
are distributed as −6 ≤ Hii ≤ 6, and therefore the diagonal disorder is smaller than
the critical disorder wc , but the off-diagonal disorder is stronger than in the calcu-
lations of Kawarabayashi et al. (1998). It has been shown (Verges, 1998; Elimes
et al., 1998) that even small off-diagonal disorder strongly enforces localization,
and thus one may conjecture that, in the 3D case, the eigenstates corresponding to
the eigenvalues n 0 are also localized for all p.
If we express the potential φ in Eq. (182) in terms of the eigenfunctions n
of H as, φ = n An n , and substitute it in Eq. (182), we obtain the following
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 111
In Eq. (185), the most important eigenstates in the sum, in the limit κ → 0, are
those with eigenvalues |m | ≤ bκ. Since the eigenvalues n are distributed in
an interval of the order of unity, the spatial density of the eigenmodes with
|m | ≤ bκ vanishes as a −d κ → 0 as κ → 0, implying that A(1) n is exponen-
0 1
tially small, |An | ∼ | m=n n |H |m Em /bm | ∝ exp −[a/ξA (0)]κ −1/d ,
(1)
and can be neglected when κ [a/ξA (0)]d . Then, the local potential φ is given
(0)
by, φ(r) = n An n = n En n (r)/(n + iκb), and the fluctuating part of
the local field Ef = −∇φ(r) is given by
Ef (r) = − En [∇n (r)/(n + iκb)] , (186)
n
is the dimensionless density of states for the Kirchhoff Hamiltonian (KH) H , and
is the system’s volume. We assume that the metal volume fraction p 1/2, so that
all quantities in the KH H are about unity, and therefore the density of states N ()
is also about unity at = 0. Hence, the distance l() can be arbitrary large as
→ 0, while it is still much smaller than the system size. It is further assumed
that the total number of eigenstates n (r) is large. When l() ξA (), the
localized eigenfunctions n (r) are characterized by spatial positions of their
centers rn , so that n (r) = (r − rn ) and Eq. (187) becomes
m
∗ ∗
n En Em [∇(1 , r − rn ) · ∇ (2 , r − rm )]
|E|2 = |E0 |2 + ,
(1 + iκb)(2 − iκb)
1 2
(189)
where the first sums are over positions of the intervals |n − 1 | and |m − 2 |
in the space, whereas the sums in the numerator are over spatial positions rn and
rm of the eigenfunctions. For each realization of a macroscopically-homogeneous
disordered film, the positions rn of the eigenfunctions (r − rn ) take on new
values that do not correlate with . Therefore, we can independently carry out the
averaging in the numerator in the second term of Eq. (189) over positions rm and
rn of eigenstates m and n . Since, ∇n (r) = 0, we obtain
∗
2 3
En Em ∇ (1 , r − rn ) · ∇ ∗ (2 , r − rm )
(190)
|En |2 |∇ (1 , r − rn )|2 δ1 2 δnm ,
The localized eigenstates are not in general degenerate, so that the eigenfunctions
n can be selected to be real, i.e., n = n∗ (where ∗ denotes the complex con-
. .2 4
jugate). Then, |En |2 = |(n |E)|2 = . N .
i=1 n,i Ei ∼ a
−2d | Edr|2 , which,
n
after using (180) and (181), yields
. .2 . .2
4−2d .
. . . .
|En | ∼ a . 4−2d . .
. n (E0 · ∇)dr. = a . (E0 · ∇n )dr. .
2
(192)
Since the local dielectric constants || are of the order of unity, one can write,
∇n ∼ n /ξA (), and therefore,
. .2 .N .2
|E0 |2 a 4 .. . . .
. ∼ |E0 | a
2 4
. .
|En | ∼
2
. n (r) dr . . n,i . . (193)
a 2d ξA2 () ξA2 () . .
i=1
Using the fact that, n |n = N
i=1 |n,i | = 1, and that n are localized within
2
ξA (), one obtains n,i ∼ [ξA ()/a] −d/2 in the localization domain which, when
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 113
and the density of states N in Eq. (197) becomes a function of p. In the limit p → 0,
the number of states effectively excited by the external field is proportional to the
number of metal particles, and hence N (p) ∼ p. The same consideration holds
in the opposite limit, p → 1, and therefore N (p) ∼ 1 − p. When N decreases,
localization becomes stronger and one can write, ξA ( = 0, p → 0) ∼ ξA ( =
0, p → 1) ∼ a. When p → 0 or p → 1, the number of the field maxima decreases
while the peaks become progressively sharper. Equation (197) also indicates that
strong field fluctuations (M2 > 1) exist in a metal-dielectric composite with d =
−m in a wide range of concentrations,
all the above results, which are based on the assumption that the eigenstates of
the Kirchhoff Hamiltonian are localized, hold in a more general case, when the
of the metal dielectric constant is negative and its absolute value is
real part m
of the order of d . The important case of |m | d will be considered in the next
subsection.
Assuming that the density of states N () and the localization length ξA () are
both smooth functions of in the vicinity of zero, and taking into account the fact
that all parameters of the Hamiltonian H for the case d = −m = 1 are of the
order of one, the following estimate for the moments of the local field is obtained
Mn,m ∼ N (p)[a/ξA (p)]2(n+m)−d κ −n−m+1 , (205)
for n + m > 1 and m > 0 (for simplicity we set b = 1). We remind the reader once
again that N (p) and ξA (p) should be understood as N (p) = N (p, = 0) and
ξA (p) = ξA (p, = 0), i.e., they are given at the eigenvalue = 0.
The maximum of the Anderson localization length ξA () is typically at the
center of the distribution of the eigenvalues (Kawarabayashi et al., 1998). When
p = 1/2, = 0 moves from the center of the -distribution toward its tails where
the localization is typically stronger (i.e., ξA is smaller). Therefore, it is plausible
that ξA (p) reaches its maximum at p = 1/2 and decreases toward p = 0 and
p = 1, so that the absolute values of the moments of the local field may have a
minimum at p = 1/2. In 2D composites the percolation threshold pc is typically
close to 0.5. Therefore, in such composites the moments Mn,m do have a local
minimum at pc as a function of the metal volume fraction p, and the amplitudes of
various nonlinear processes, while much enhanced, have a characteristic minimum
at pc . It is important to note that the magnitude of the moments in Eq. (205) do not
depend on the number of annihilated photons in one elementary act of the nonlinear
scattering. However, when all photons are added (i.e., when all frequencies enter
the nonlinear susceptibility with the plus sign) and n = 0, one cannot estimate the
moments M0,m by Eq. (205), since the integral in Eq. (204) is no longer determined
by the poles at = ±ibκ. However, all the functions of the integrand are about
unity and M0,m ∼ O(1) for m > 1. The moment M0,m is an important quantity
since it yields the enhancement GnH G of the nth order harmonic generation through
the relation, GnH G = |M0,m |2 (see below).
116 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
account the fact that the electric field renormalizes as E0 = E0 (br /a), one obtains
from Eq. (198) the following expression for the field’s peaks in the renormalized
system:
$ % $ %
|m | ν/(µ+s) |m |
Em E0 (a/ξA )2 (br /a)κ −1 E0 (a/ξA )2
, (207)
d m
where ξA = ξA (pc ) is the localization length in the renormalized system. Each
maximum of the field in the renormalized system is in a dielectric gap in a dielectric
cube of linear size br or in between two conducting cells of the size br that are
not necessarily connected to each other. There is not a characteristic length in the
original system which is smaller than br , except the grain size a. Therefore, it is
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 117
plausible that the width of a peak of the local field in the original system is about
a. Then, values of the field maxima Em do not change when returning from the
renormalized system to the original one. Hence, Eq. (207) yields values of the field
maxima in the original system.
Equation (207) provides the estimate for the local field extrema when the real part
of the dielectric constant is negative. For metals increases in absolute value
m m
with the wavelength, when the frequency ω < ω̃p . Therefore, the field maxima
Em (ω) increase strongly with the wavelength. For a Drude metal the steep growth
of the peaks Em (ω) occurs for the frequencies ω < ω̃p , when the dielectric constant
m can be approximated as
b b ωτ
m (ω < ω̃p ) 2(ω − ω̃p ) +i , (208)
ω̃p ω̃p
which, when substituted in Eq. (207), yields
. . (ν+µ+s)/(µ+s)
. .
2 2b ω − ω̃p ω̃p
Em (ω < ω̃p ) E0 (a/ξA ) ν/(µ+s)
.
ω̃p ωτ b d
(209)
Since in a typical metal, ωτ ω̃p , the amplitudes of the field’s peak first increase
steeply and then saturate (see below) at Em E0 (a/ξA )2 (b /d )ν/(µ+s) (ω̃p /ωτ ) ∼
E0 ω̃p /ωτ , when ω 0.5ω̃p . Therefore, the intensity maxima Im exceed the
2
intensity of the incident wave I0 by a factor Im /I0 ∼ ω̃p /ωτ 1.
Consider now the case ω ωp , when for a Drude metal
$ %2
ω ωτ
m (ω ωp ) − 1−i , ω ωτ (210)
ωp ω
which, when substituted in Eq. (207), yields
$ %2 $ % $ %
a ωp 2ν/(µ+s) ω
Em (ω ωp ) E0 √ . (211)
ξA d ω ωτ
For 2D percolation, the critical exponents are, µ = s ν = 4/3, √and thus
√
Eq. (211) yields, Em ∼ E0 (a/ξA )2 ωp /( d ωτ ) = E0 (a/ξA )2 (ω̃p /ωτ ) b /d ∼
E0 (ω̃p /ω), which coincides with the estimate obtained from Eq. (209) for ω =
0.5ω̃p , implying that the local field’s peaks increase steeply when m , the real
part of m , is negative and then remains essentially constant in the wide frequency
range, ω̃p < ω < ωτ .
For 3D percolation composites, we roughly have, ν (µ + s)/3, and thus
2/3
Eq. (211) yields, Em ∼ E0 (b /d )1/3 ω̃p ω1/3 /ωτ , implying that the local field
peaks increase up to Em /E0 ∼ ω̃p /ωτ when m < 0, and then decrease as
To obtain Mn,m we consider first the spatial distribution of the maxima of the
field for |m | d . The average distance between the maxima in the renormalized
system is ξe , given by Eq. (199). Then, the average distance ξe between the maxima
118 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
localization length in the original system, i.e., a typical size of the eigenfunction,
is about (br /a)ξA a, i.e., the eigenstates become macroscopically large when
|m | /d 1, and consist of sharp peaks separated in space by distances much
larger than a.
It is then not difficult to show, using Eq. (215), that
$ % $ %
m n(br ) m |m | (m−2+s/ν)ν/(µ+s)
M0,m ∼ M0,m (br /a)
∼ , (216)
(ξe /a)d |m | d
which holds when M0,m > 1. In 2D, if we use µ = s ν = 4/3, Eqs. (215) and
(216) are simplified to
n+m−1
|m |3/2
Mn,m ∼ N ! , (217)
(ξA /a)2 d m
for n + m > 1 and n > 0, and
| |(m−3)/2
m m
M0,m ∼ (m−1)/2
, (218)
d
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 119
. The moments M
for m > 1, n = 0 and (m |/d )(m−1)/2 > |m |/m n,m (n = 0) are
strongly enhanced in 2D Drude metal-dielectric composites since they reach the
maximum value
n+m−1
ωp
Mn,m ∼ N √ , (219)
ωτ d (ξA /a)2
when ω ωp . Thus, in a 2D percolation composite the moments Mn,m are
independent of frequency if ω ωp . For metals this typically takes place in
the red and infrared spectral ranges. For example, for a semi-continuous sil-
ver film on a glass substrate, the moments Mn,m can be estimated as, Mn,m ∼
[3 × 102 (a/ξA )2 ]n+m−1 , for ω ωp .
It follows from Eq. (215) that for 3D metal-dielectric percolation composites,
for which the dielectric constant of the metal component can be estimated by the
Drude formula, the moments Mn,m (n = 0) achieve their maximum at frequency
ωmax 0.5ω̃p . Since, as mentioned above, for 3D percolation, ν/(µ + s) 1/3,
the maximum value of Mn,m is roughly given by
n+m−1
Mn,m (ω = ωmax ) ∼ N (ξA /a) (a/ξA )2 (b /d )1/3 ω̃p /ωτ , (220)
whereas for ω ωp ,
2/3
n+m−1
(a/ξA )2 ωp ω1/3
Mn,m ω ωp ∼ N (ξA /a) 1/3
. (221)
d ωτ
Figure 3.10 compares the results of numerical and theoretical calculations for
Figure 3.10. Moments Mn,m of the electric field in semicontinuous silver films versus the
wavelength λ at the percolation threshold. On the left are the moments Mn = Mn,0 , from
the bottom to the top, for n = 2, 3, 4, 5 and 6. The solid curves are the predictions of the
scaling theory, Eq. (215), while the symbols are the numerical simulation data. Shown on
the right are the moments M4,0 (upper solid curve predicted by the scaling theory, versus ∗,
the numerical data), M0,4 (upper dashed curve), M2,0 (lower solid curve predicted by the
scaling theory, versus +, the numerical data), and M0,2 (lower dashed curve predicted by
the scaling theory, versus circles, the numerical data) (after Sarychev and Shalaev, 2000).
120 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
analysis was based on the finite-size scaling analysis (see Chapter 2 of Volume I for
description of the finite-size scaling method), which holds provided that lr < ξp ,
where ξp is the percolation correlation length. Since at pc the correlation length
ξp diverges, these estimates are valid in the wide frequency range ωτ < ω < ω̃p ,
which includes the visible, infrared, and far-infrared spectral ranges for typical
metals. For any particular frequency from this interval, one can estimate the range
p, where the giant field fluctuations occur, by requiring that, lr = ξp , which
results in, |p| ≤ (d / |m |)1/(µ+s) . Therefore, the local electric field fluctuates
strongly for such volume fractions and its moments Mn,m are much enhanced.
In the discussion that follows, the skin effect is neglected so that a semi-
continuous film can be considered as a 2D material. In the optical frequency range,
where the frequency ω is much larger than the relaxation rate τ −1 of the metallic
component, a semi-continuous metal film can be thought of as a 2D L − R − C
lattice (see, for example, Brouers et al., 1993). As before, the capacitance C rep-
resents the gaps between metal grains that are filled by the dielectric material
(substrate) with a dielectric constant d . The inductive elements L − R represent
the metallic grains that for the Drude metal have the dielectric function m (ω)
given by Eq. (177). In the high-frequency range considered here, the losses in the
metal grains are small, ω ωτ . Therefore, m (in modulus) and < 0
√ m m
for frequencies ω < ω̃p = ωp / b . Thus, the metal conductivity is almost purely
imaginary and the metal grains can be modeled as L-R elements, with the active
component being much smaller than the reactive one. If the skin effect cannot be
neglected, i.e., if the skin depth δ < a, the simple quasi-static presentation of a
semi-continuous film as a 2D array of the L − R and C elements is not valid. One
can still use the L − R − C model in the other limiting case, when the skin effect
is very strong (δ a). In this case, the losses in the metal grains are small, regard-
less of value of ω/ωτ , whereas the effective inductance for a metal grain depends
on the grain size and shape rather than on the material constants for the metal.
It is instructive to consider first the properties of the film at p = pc , where
the duality relation (see above and also Chapters 4 and 5 of Volume I) predicts
that, the effective dielectric constant e in the quasi-static case is given exactly by,
√
e = d m . If we neglect the metal losses and set ωτ = 0, the metal dielectric
constant m < 0 for ω < ω̃p . We also neglect possible small losses in a dielectric
substrate, assuming that d is real and positive, in which case e is purely imaginary
for ω < ω̃p . Therefore, a film consisting of loss-free metal and dielectric grains
is absorptive for ω < ω̃p . The effective absorption in a loss-free film means that
the electromagnetic energy is stored in the system and thus the local fields could
increase without limit. In reality, due to losses the local fields in a metal film are,
of course, finite. However, if the losses are small, one may expect very strong
fluctuations in the field. To calculate Rayleigh and Raman scattering, and various
nonlinear effects in a semi-continuous metal film, one must know the field and
current distributions in the film.
Although, as discussed in Chapters 4 and 5 of Volume I, there are several very
efficient numerical methods for calculating the effective conductivity of composite
materials, they typically do not allow calculations of the field distributions. Brouers
et al. (1997) developed a PSRG method, a generalization of what was described
in Chapter 5 of Volume I, using a square lattice of the L − R (metal) and C
(dielectric) bonds. A fraction p of the bonds were metallic (L − R bonds) and had
a conductivity gm = −im ω/4π, while the dielectric (C) bonds, with a fraction
1 − p, had a conductivity gd = −id ω/4π. The applied field E0 was E0 = 1,
whereas the local fields inside the system were of course complex quantities. In
this method, after each RG transformation, an external field E0 is applied to the
system and the Kirchhoff’s equations are solved in order to determine the fields
and the currents in all the bonds of the transformed lattice. The self-dual PSRG cell
122 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
F F
F F
F
F F
F F F
F
F F
of Figure 3.11 was used which, because of its hierarchical structure, allows these
equations to be solved exactly. Then, the one-to-one correspondence between the
elementary bonds of the transformed lattice and the bonds of the initial square
lattice was used for determining the field distributions, as well as the effective
conductivity, of the initial lattice. The number of operations for obtaining the full
distributions of the local fields is proportional to b2 [to be compared with O(b7 )
operations needed in the transform-matrix method and O(b3 ) operations needed
in the Lobb-Frank algorithm that was described in Sections 5.14.2 and 5.14.3 of
Volume I]. The Drude formula for metal dielectric functions was used, and thin
films of silver (for which b = 5, the plasma frequency ωp = 9.1 eV, and the
relaxation frequency ωτ = 0.021 eV) and gold (for which b = 6.5, ωp = 9.3 eV,
and ωτ = 0.03 eV), deposited on a glass substrate with the dielectric constant
d = 2.2, were modeled.
All the numerical results obtained with this method were in agreement with the
predictions of the scaling theory discussed above, as well as with experimental
data, described below.
The metal-insulator transition is very close to this point, even in the presence of
quantum tunneling. At higher surface coverage, the film is mostly metallic, with
voids of irregular shapes. As coverage increases further, the film becomes uniform.
Optical properties of such metal-dielectric films exhibit anomalous phenomena that
are absent for bulk metal and dielectric components. For example, the anomalous
absorption in the near-infrared spectral range leads to unusual behavior of the
transmittance and reflectance in that, the transmittance is much higher than that of
continuous metal films, whereas the reflectance is much lower.
The predictions of the PSRG computations have been compared with the experi-
mental data for gold-on-glass films at various wavelengths (Sarychev and Shalaev,
2000). There is good qualitative agreement between the two. The data for such dis-
ordered metal-dielectric films near pc suggest localization of optical excitations
in small nm-scale hot spots. The hot spots of a percolation film represent very
large local fields (fluctuations); spatial positions of the spots strongly depend on
the light frequency. Near-field spectra observed and calculated at various points of
the surface consist of several spectral resonances, the spectral locations of which
depend on the probed site of the sample. These features are observable only in
the near zone. In the far zone, one observes images and spectra in which the hot
spots and the spectral resonances are averaged out. The local field enhancement is
large, which is especially important for nonlinear processes of the nth order, and
are proportional to the enhanced local fields to the nth power. This opens up a fas-
cinating possibility for nonlinear near-field spectroscopy of single nano-particles
and molecules.
above to carry out the analysis, which yields interesting results. For example, using
a system of size b = 1024, p = pc = 0.5, and ω = ωr , Sarychev and Shalaev
(2000) calculated the electric field in all the bonds for 10−4 ≤ κ ≤ 10−1 with the
external field being E0 = 1. The distribution of the field intensity, I (r) = |E(r)|2 ,
was found to be close to the well-known log-normal distribution, with its values
spread over many orders of magnitude, even for κ = 10−1 . For κ = 10−4 , I (r)
was distributed essentially uniformly in (0, 104 ). The average intensity, I =
|E0 |2 M2 , increased as, I ∝ κ −1 , in agreement with Eq. (205). Thus, the field
fluctuations lead to enhanced light scattering from the film.
It should be pointed out that the fluctuations considered here, and the corre-
sponding light scattering, do not arise because of the fractal morphology of the
metal clusters, but are due to the distribution of local resonances in a disordered
metal-dielectric film, which is homogeneous on a macroscopic scale. The local
intensity of the electric field is strongly correlated in space, and the distribution is
dominated by the field correlation length ξe introduced by Eqs. (199) and (212),
and defined as the length scale over which the field fluctuations are small. As the
L− (metallic) component becomes loss-free (κ → 0), ξe diverges according to
ξe ∼ κ −νe , (223)
where νe is a new critical exponent which has been estimated by several numerical
methods. For example, in 2D the PSRG method described above yields νe
0.45 ± 0.05, while the scaling theory, Eq. (208), predicts that νe = 1/d, where d
is the space dimension, a result that was also conjectured by Hesselbo (1994). For
small loses at resonance, the correlation length ξe is the only relevant length scale
of the system at pc since |m |/d 1.
for r ∼ a is given by, C(0) ∼ C(a) ∼ M2 (lr /a)µ/ν ∼ M2 (|m |/d )µ/(s+µ) , and
hence
$ %
|m | (2s+µ)/(s+µ)
Mj ∼ (ω/4π )2 |E0 |2 d2 M2 ∼ (ω/4π )2 |E0 |2 d |m | M2 ,
d
(238)
where M2 ≡ M2,0 and M ≡ M , as defined earlier. Equation (238) holds for
2 2,0
arbitrary spatial dimension.
We now consider light scattering from semi-continuous metal films for ω
ωp , where the metal dielectric constant for a Drude metal is approximated as,
m −(ωp /ω)2 (1 − iωτ /ω). By substituting Eq. (237) into (233) and taking into
account the fact that at pc , |e |2 d |m | d (ωp /ω)2 (using µ = s ν = 4/3),
the following result is obtained
(ka)4
Itot ∼ |T |2 |e |2 κ −1 br 1+η ξe 1−η
3
(ka)4 ω a 4 $ ω %1+(1−η)/2
2 −1−(1−η)/2 p
∼ |T | κ |m | ∼ 0.1
2
, (239)
3 c ωτ
where the experimental result, |T |2 0.25, which holds for p = pc and ωτ
ω ωp , was used. Thus, the scattering first increases as ω1+(1−η)/2 with increas-
ing ω according to Eq. (239) and then vanishes as (ω̃p − ω)2+(1−η)/2 as ω → ω̃p
[see Eq. (236)].
The enhancement of the scattering due to the field fluctuations can be estimated
from Eqs. (234) and (237) as, Ig ∼ |T |2 d |m |br2 κ −1−(1−η)/2 /4, which yields, for
a Drude metal and ω ωp , the following equation
$ % $ %
|T |2 ω̃p 4 ω 1+(1−η)/2
Ig ∼ . (240)
4 ω ωτ
Using typical values, |T |2 = 1/4 and d = 2.2, the enhancement Ig can become as
large as 5 × 104 at wavelength λ = 1.5 µm and continues to increase towards the
far infrared spectral range. Note that Rayleigh scattering decreases as ω4 with de-
creasing frequency, whereas the anomalous scattering varies as, I ∼ ω1+(1−η)/2
ω1.1 , and therefore the enhancement increases as Ig ∼ ω−2.9 ∼ λ2.9 in the infrared
part of the spectrum.
roughness features of various shapes, and (2) surface plasmon waves (polaritons)
that laterally propagate along the metal surface. In practice, there are strong light-
induced interactions between different features of a rough surface, and therefore
plasmon oscillations should be treated as collective surface excitations (localized
surface plasmons) that depend strongly on the surface morphology.
field E0 is constant in the film plane. The local field E(r2 ), induced by the external
field E0 , is obtained by using Eq. (144) for the non-local conductivity ,
1
E(r2 ) = (r2 , r1 )E0 dr1 , (246)
g(r2 )
and excite Raman-active molecules that are (assumed to be) uniformly distributed
in the composite. Such molecules, in turn, generate the Stokes fields, Es (r2 ) =
αs (r2 )E(r2 ), oscillating at the shifted frequency ωs , where αs (r2 ) is the ratio of
the Raman and linear polarizabilities of the Raman-active molecules at r2 . The
Stokes fields Es (r2 ) induce in the composite the currents Is (r3 ) that are given by
Is (r3 ) = (r3 , r2 )Es (r2 ) dr2 . (247)
Since the frequency ωs is typically close to the external field’s frequency, i.e.,
|ω − ωs |/ω 1, the non-local conductivities appearing in Eqs. (246) and (247)
are essentially the same.
The intensity I of the electromagnetic wave scattered from any inhomogeneous
material is proportional to the current fluctuations inside the system:
,. .2 -
. .
I ∝ .. [I(r) − I] dr.. , (248)
where the integration is over the entire system, and · denotes an ensemble average.
For Raman scattering, · also includes averaging over the fluctuating phases of
the incoherent Stokes fields generated by Raman-active molecules. Therefore, the
average current densities oscillating at ωs is zero, Is = 0, and hence the intensity
IR of Raman scattering from a semi-continuous metal film is given by
, . .2 -
. .
IR ∝ . I(r) dr..
.
∗ ∗ ∗
= αβ (r3 , r2 )αs (r2 )Eβ (r2 ) αγ (r5 , r4 )αs (r4 )Eγ (r4 ) dr2 dr3 dr4 dr5
(249)
where a summation over repeating Greek indices is implied, and the integration
is over the entire film plane. Equation (249) is now averaged over the fluctuating
phases of the Raman polarizabilities αs . Because the Raman field sources are
incoherent, we have αs (r2 )αs∗ (r4 ) = |αs |2 δ(r2 − r4 ), and therefore
∗ ∗
IR ∝ αβ (r3 , r2 ) µγ (r5 , r2 )δαµ |αs | Eβ (r2 )Eγ (r2 ) dr2 dr3 dr5 . (250)
2
If we now take advantage of the facts that, (1) a semi-continuous film is macroscop-
ically homogeneous, and thus its Raman scattering is independent of the orientation
of the external field E0 ; (2) due to (1), Eq. (250) can be averaged over the orien-
tations of E0 without changing the result, and (3) the non-local conductivity is
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 131
3/2
that, IRS ∼ N (p)[a/ξA (p)]6 (ωp /ωτ )3 /d , independent of the frequency. For
example, for silver-on-glass percolation films at pc , the Anderson localization
length ξA is about ξA 2a, the density of state, N (pc ) 1, and therefore, IRS ∼
106 . For 3D composites at ω ωp , IRS decreases with decreasing ω as IRS ∼
N (p)(ξA /a)−5 ωp2 ω/ωτ3 ∼ 106 ω/ωp , where the 3D critical exponents have been
approximated as, ν s (µ + s)/3, and the data, ωp = 9.1 eV and ωτ = 0.021
eV, for silver dielectric constant have been utilized.
Consider now hyper-Raman scattering when n photons of frequency ω are con-
verted to one hyper-Stokes photon of the frequency ωhRS = nω − ωsf , where
ωsf is the Stokes frequency shift corresponding to the frequency of molecule os-
cillations (electronic or vibrational). Thus, following the same line of reasoning
outlined above, the surface enhancement of hyper-Raman scattering (SEHRS)
IhRS is given by
|ghRS (r)|2 |EhRS (r)|2 |E(r)|2n |hRS (r)|2 |EhRS (r)|2 |E(r)|2n
IhRS = . .2 = . .2 ,
|gd |2 .E0,hRS . |E0 |2n |d |2 .E0,hRS . |E0 |2n
(256)
where EhRS (r) is the local field excited in the system by the uniform probe
field E0,hRS , oscillating with ωhRS , and ghRS (r) and hRS (r) are the local con-
ductivity and dielectric constant at frequency ωhRS . For n = 1 Eq. (256) describes
the conventional SERS. To estimate IhRS , we must keep in mind that the spatial
scales br for the field maxima at the fundamental frequency ω and the hyper-Stokes
frequency ωhRS are significantly different. Therefore, the average in Eq. (256) can
be decoupled and approximated as, |hRS (r)|2 |EhRS (r)|2 |E(r)|2n ∼ |hRS (r)
EhRS (r)|2 |E(r)|2n = |hRS (r)EhRS (r)|2 M2n |E0 |2n , where M2n (ω) is the
2nth moment. It follows from Eq. (238) that, |hRS (r)EhRS (r)|2 ∼ d |m (ωhRS )
|M2 |E0,hRS |2 , where M2 (ωhRS ) is the second moment of the field EhRS (r). Using
the expressions for M2 and M2n given above, and taking into account the fact
that for p pc the density of states N is about unity, one obtains the following
equation for enhancement of hyper-Raman scattering,
$ % $ %
|m (ωhRS )| (µ+2s)/(µ+s) |m (ωhRS )|
IhRS ∼ (ξA /a)2d−4(1+n) (ω
d m hRS )
$ %[2ν(n−1)+s]/(µ+s) $ %2n−1
|m (ω) | |m (ω)|
× (ω)
, (257)
d m
with n ≥ 2. For a Drude metal and frequencies ω ω̃p , ωhRS ω̃p the metal
dielectric constant can be approximated as, |m (ωhRS )| ∼ |m (ω)| ∼ (ωp /ω)2 and
(ω)/| (ω)| ∼ ω /ω, and therefore Eq. (257) becomes
m m τ
2[2ν(n−1)+3s+µ]/(µ+s) $ ω %2n
2d−4(1+n) ωp
IhRS ∼ (ξA /a) , (258)
ω ωτ
which, in 2D (using µ = s ν = 4/3), simplifies to
2(n+1) $ ω %2n 2 $ ω %2n
4n ωp 4n ωp p
IhRS ∼ (a/ξA ) ∼ (a/ξA ) . (259)
ω ωτ ω ωτ
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 133
Figure 3.12. Comparison of experimental data (points with error bars) for normalized
SERS, I¯ = IRS (p)/IRS (p = pc ), for a semicontinuous silver film, versus the theoretical
computations (curve) (after Sarychev and Shalaev, 2000).
where |E0 | is the amplitude of the external electric field at frequency ω, and α and
β are some constants. Note that, for an isotropic film, the second term in Eq. (261)
results in change of the polarization of the incident light. Moreover, for the case
of linear and circular polarization of the incident light, Eq. (261) can be simplified
since for linear polarization the complex vector E0 reduces to a real vector. Then,
|E0 |2 E0 = E02 E0 , and Eq. (261) becomes
(3)
where the effective nonlinear dielectric constant e is now a scalar quantity. Let
us consider, for the sake of simplicity, the linearly polarized incident wave. We
write Eq. (262) in terms of the nonlinear average current I(3) (r) and the effective
(3) (3)
Kerr conductivity ge = −iωe /4π:
(3)
We consider first the limit in which the nonlinearities in metal grains gm and
(3) (3) (3)
dielectric gd are approximately equal, gm gd , which can be caused by, for
example, molecules that are uniformly covering a semi-continuous film. Then
. .2
I(r) = g () (r)E (r) + g (3) .E (r). E (r), (264)
where E (r) is the local fluctuating field. Then, current conservation law takes the
following form
g (3) .. ..2
∇ · g (r) −∇φ(r) + E0 + () E (r) E (r)
()
= 0, (265)
g (r)
where −∇φ(r) + E0 = E (r) is the local field. The second and third terms of
Eq. (265) can be thought of as a renormalized external field
g (3) .. ..2
Ee (r) = E0 + Ef (r) = E0 + E (r) E (r) , (266)
g () (r)
where the field Ef (r) may change over the film but its average, Ef (r), is collinear
to E0 , in which case the average current density I(r) is also collinear to E0 and
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 135
can be written as
E0 E0 1
I = (E0 · I) = 2 E0 · I(r) dr, (267)
E02 E0 A
where A is the total area of the film, the integration is over the film area, and
E02 ≡ (E0 · E0 ). Expressing I(r) in terms of the non-local conductivity matrix
defined by Eq. (241) yields
E0 1
I = 2 [E0 (r, r1 )Ee (r1 )] dr dr1 . (268)
E0 A
If we integrate Eq. (268) over the coordinates r and use the symmetry of the
non-local conductivity matrix , we obtain
E0 1
I = 2 [I0 r · Ee (r)] dr, (269)
E0 A
where I0 (r) is the current induced at r by the constant external field E0 . Using
Eq. (266) and carrying out the integration, Eq. (269) becomes
⎡ 2 3. .2 ⎤
g (3) E(r) · E (r) .E (r).
I = E0 ⎣ge() + ⎦, (270)
E02
()
where ge and E(r) are the effective conductivity and local fluctuating field in
the linear approximation (i.e., for g (3) ≡ 0). Comparison of Eqs. (270) and (263)
yields an expression for the effective Kerr conductivity:
2 3. .2
g (3) E(r) · E (r) .E (r).
ge(3) = . (271)
E02 |E0 |2
Equation (271) is general and applicable to weak as well as strong nonlinearities.
In the former case, E (r) E(r), and Eq. (271) becomes
g (3) E 2 (r)|E(r)|2
ge(3) = , (272)
E02 |E0 |2
(3)
yielding ge in terms of the linear local field. Note that Eq. (272) is the analogue
(3)
of (128). In the absence of metal grains, ge = g (3) . Therefore, the enhancement
IK of the Kerr nonlinearity is given by
E 2 (r) |E(r)|2
IK = = M2,2 , (273)
E02 |E0 |2
where M2,2 is the fourth moment of the local field.
(3) (3) (3)
Equations (272) and (273) were derived assuming that gm gd . If gm =
(3)
gd , the above analysis can be repeated in order to derive the following equation,
E 2 (r) |E(r)|2 (3) E 2 (r)|E(r)|2
ge(3) = pgm
(3) m
+ (1 − p)gd d
, (274)
E02 |E0 |2 E02 |E0 |2
136 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
where ·m and ·d represent averaging over the metal and dielectric grains, re-
spectively. Note that, for the case of cubic nonlinearity in the conductivity of
materials with percolation disorder, Eq. (274) was already derived and discussed
in Section 3.2 [see Eq. (128)]. For the case of Kerr conductivity, Eq. (274) was first
derived by Shalaev et al. (1998). According to Eq. (273), the Kerr enhancement
IK is proportional to the fourth power of the local field, averaged over the sample,
which is similar to the case of SERS with the enhancement factor IRS given by
Eq. (254). Note, however, that while IK is complex, IRS is a real and positive
quantity.
The enhancement of the Kerr nonlinearity can be estimated analytically using
the methods described above. Consider first the case when g (3) (r) in the dielectric
component is of the same order of magnitude or larger than in the metal component.
Then,
. . . .
. . . .
IK ∼ .ge(3) / g (3) (r) . = .e(3) / (3) (r) . ∼ |M2,2 |
$ % $ %
|m | (2ν+s)/(µ+s) |m | 3
∼ N (ξA /a)d−8
, (275)
d m
where Eq. (215) has been used for the moment M2,2 . For ω ωp , the Kerr en-
hancement for 2D composites is estimated as, IK ∼ N (ξA /a)−6 (ωp /ωτ )3 , if the
Drude formula is used for the metal dielectric constant m . For example, as dis-
cussed above, for silver-on-glass semi-continuous films, Anderson localization
length ξA 2a and density of states, N 1, and therefore, IK ∼ 105 − 106 . As
discussed by Sarychev and Shalaev (2000), for d = 2 a plot of IK versus the metal
volume fraction p has a two-peak structure, which is similar to the case of Raman
scattering shown in Figure 3.12. However, in contrast to IRS , the dip at p = pc
is much more pronouced and is proportional to the loss factor κ, implying that at
p = pc the enhancement is actually given by, IK ∼ κM2,2 . This result is presum-
ably a consequence of the special symmetry of a 2D self-dual system at p = pc .
If one .moves. slightly away from p = pc , the enhancement IK increases such that,
IK ∼ .M2,2 . ∼ IRS ∼ M4,0 . The fact that the minimum at p = pc is much smaller
for SERS than for the Kerr process is presumably related to the latter being a phase
sensitive effect. Moreover, as already discussed above, the local field maxima are
concentrated in the dielectric gaps where |m | d . Therefore, Eq. (275) is valid
when the Kerr nonlinearity is located mainly in such gaps.
Consider now the case when the Kerr nonlinearity is due to metal grains (see,
for example, Ma et al., 1998; Liao et al., 1998). Provided that m − , the
d
local electric fields are equally distributed in the metal and dielectric components,
implying that the Kerr enhancement is still given by Eq. (274) with |m |/d = 1.
However, if m | d , the local field will be concentrated in the dielectric space
between the conducting clusters with a value Em given by Eq. (207). The total
current Is of the electric displacement flowing in the dielectric space between
two resonate metal clusters of size br is given by, Is = aEm e brd−2 . Because of
the current continuity, the same current should flow in the adjacent metal clusters
where it is concentrated in a percolating channel. The electric field Emc in the metal
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 137
channel, which spans over the cluster, is given by, Emc ∼ Is /(m a d−1 ), where
a d−1 represents the cross-section of the channel. Then the nth moment of the local
electric field in a metal cluster of size br is, Emc
n = E n La d−1 /bd , where L =
mc r
−d+2
a(m /e )br is the effective length of the conducting channel. Keeping in mind
that only a fraction κ = m /| | 1 of the metal clusters of size b are excited
m r
by the external electric field, we obtain, Mnmet = |E|n met /E0n = κEmc n /E n , for
0
the moments of the electric field in the metal component,
$ % $ %
|m | n−1 |m | [(d−1)(n−2)ν−µ(n−1)]/(µ+s)
Mn ∼
met
, (276)
m d
where Eq. (206) was used for the size br of the resonant clusters. Then,
enhancement IKmet of the Kerr nonlinearity is given by
$ % $ %
|m | 3 |m | [2(d−1)ν−3µ]/(µ+s)
IKmet ∼ M4met ∼
. (277)
m d
)3
In 2D (for which µ = s ν = 4/3) Eq. (277) yields, IKmet ∼ M4met ∼ (|m | /m
(d / |m |) . As expected, IK IK , and in fact for 2D systems near pc ,
1/2 met
$ %
IK |m | 2
∼ . (278)
IKmet d
Since in optic and infrared spectral ranges, |m | d , the enhancement due to
the Kerr nonlinearity is much larger than when the initial nonlinearity is located
in the dielectric gaps where the local fields are much larger than in the metal. It
follows from Eq. (278) that the Kerr enhancement IKmet may become less than one,
implying that, on average, the local electric field in the metal component can be
smaller than the external field. For example, for semi-continuous silver films on a
glass substrate, IKmet < 1 for wavelengths λ > 10µm.
covered by a layer possessing a nonlinear conductivity g (n) . The layer can be made
of nonlinear organic molecules, semi-conductor quantum dots, or a quantum well
on top of a percolation film. The local electric field Eω (r), induced in the film by
the external field E0 , generates in the layer the nω current g (n) Eω Eωn−1 . Note that,
strictly speaking, this expression is valid only for the scalar nonlinear conductiv-
ity and odd n. However, for obtaining order-of-magnitude estimates, we can use
this formula for arbitrary n. The external field, oscillating at frequency ω, is still
denoted as E0 , though the frequency is indicated explicitly for other fields. The
nonlinear current g (n) Eω Eωn−1 , in turn, interacts with the film and generates the
initial nω electric field with an amplitude E(n) = g (n) Eωn−1 Eω /g () , where g () is
the linear conductivity of the nonlinear layer at frequency nω. The electric field
E(n) can be thought of as an inhomogeneous external field exciting the film at
frequency nω.
The nHG current I(n) induced in the film by the initial field E(n) can be de-
termined in terms of the non-local conductivity matrix (r, r ) introduced by
Eq. (241):
Iβ (r) = βα (r, r )Eα(n) (r ) dr ,
(n) (n)
(279)
(n)
where βα is the conductivity matrix at frequency nω, the integration is over the
entire film area, the Greek indices represent {x, y}, and summation over repeated
indices is implied. It is I(n) that eventually generates the nonlinear scattered field at
frequency nω. By using the standard approach of the scattering theory adopted to
semi-continuous metal films (Brouers et al., 1998), and assuming that the incident
light is unpolarized, the integral scattering in all directions but the specular one is
given by
$ . . %
4k 2 . (n) .2
I= Iα (r1 )Iα (r2 ) − . I .
(n) (n)∗
dr1 dr2 , (280)
3c
where the integrations is over the entire area A of the film, k = ω/c, and ·
indicates an ensemble average. As in the case of Rayleigh scattering, we have
assumed that the integrand vanishes for r λ, where r = r2 − r1 [therefore, the
term exp(ik · r) was omitted]. Using Eq. (279), we can write
Iα(n) (r1 ) Iα(n)∗ (r2 ) dr1 dr2
'
4
Eβ (r3 )Eα∗(n) (r4 )
(n) (n)∗ (n)
= γβ (r1 , r3 ) δα (r2 , r4 )δγ δ dri , (281)
0
i=1
where ·0 denotes an average over the light polarization. We now introduce the
(0)
spatially uniform probe field Enω which oscillates at frequency nω and is assumed
(0) (0)∗ (0)
to be unpolarized. For the unpolarized light, δγ δ = 2 Enω,gamma Enω,δ /|Enω |2 ,
0
which, when substituted in Eq. (281), the integration is carried out over the coor-
(0)
dinates r1 and r2 , and the averaging over independent polarizations of fields Enω
and E0 are performed, the following equation for the current-current correlation
3.5. Electromagnetic Field Fluctuations and Optical Nonlinearities 139
function is obtained,
Iα(n) (r1 ) Iα(n)∗ (r2 ) dr1 dr2 =
1 ∗
2 3
(0)
()
gnω (r3 )gnω (r4 ) Enω (r3 ) · E∗nω (r4 ) E(n) (r3 ) · E(n)∗ (r4 ) dr3 dr4 ,
|Enω |2
(282)
(0)
where Enω (r) is the local nω field excited in the film by the probe field Enω , and
()
gnω (r) is the film linear conductivity at frequency nω.
In macroscopically-homogeneous and isotropic films considered here, the inte-
(0)
gral in Eq. (282) does not depend on direction of the probe field Enω . Therefore,
(0)
Enω can be selected to be collinear with the external field E0 . Moreover, I(n) is
(0)
parallel to the external field E0 . If the probe field Enω is aligned with E0 , we have,
.
. (n) .2 . (0) (n) .2 .
. I . = . Enω · I . /|E(0) nω | . Then, using Eq. (279), we can write
2
. . . .2
. (n) .2 1 . .
.I . = . E (0) (n)
(r , r )E (n)
(r ) dr dr .
2. . (283)
(0) 2 . nω,β βα 1 2 α 2 1
A|Enω |
If the integration over coordinate r1 is carried out, one obtains
. ..2
. .2 . ()
. gnω Enω · E(n) .
. (n) .
.I . = (0)
, (284)
|Enω |2
and therefore
. .2 5 6 ∞
8π k 2 .. g (n) .. .
. () .2
.
I= . () . A .gEnω . |Eω | |Eω | 2 2(n−1)
C (n) (r)r dr, (285)
3c|Enω |2 . g .
(0)
0
where it was assumed that the generated frequency nω is less than ωp , so that
3.6. Electromagnetic Properties of Solid Composites 141
current. Therefore, the inside electric field consists of uniform curl-free part Ein,m0
(i.e., ∇ × Ein,m0 ) and the rotational part L(r) that depends on the coordinate. The
field outside the metal particle is given by
$ %
e − ˜m E0 · r
Eout,m = E0 + a 3 ∇ . (296)
2e + ˜m r3
√
The local wavelength inside a dielectric grain, λd = λ/ d , is assumed to be much
larger than the grain size a. Then, the electric fields inside and outside a dielectric
particle are given by the following well-known equations, already familiar from
Chapters 4 and 5 of Volume I:
3e
Ein,d = E0 , (297)
2e + d
$ %
3 e − d E0 · r
Eout,d = E0 + a ∇ . (298)
2e + d r3
Similar equations can be obtained for the magnetic field excited by a uniform
magnetic field H0 inside and outside a metal (dielectric) particle:
Hin,m = Hin,m0 + 4πM, (299)
where
3Ke
Hin,m0 = H0 , (300)
2Ke + K̃m
and the renormalized metal magnetic permeability K̃m is given by
2F (ym a)
K̃m = Km , (301)
1 − F (ym a)
where the function F is defined by Eq. (293). Note that the renormalized metal
magnetic permeability K̃m is not equal to one, even if the metal is non-magnetic and
the seed magnetic permeability Km = 1. The local magnetic field M in Eq. (299)
is the solution of
1 ik
∇×M = ∇ × Hin,m = − DH , (302)
4π 4π
with
akKm Ke sin(ym r)F (km r) 7y x 8
DH = 3iH0 ,− ,0 , (303)
(2Ke + K̃m ) sin(ym a)[F (ym a) − 1] r r
where DH is the electric displacement induced in the metal particle by high-
frequency magnetic field H0 . The displacement DH can be written as DH =
i(4π/ω)I, where the eddy electric current I is called the Foucault current. The
field Hin,m0 is the potential (curl-free) part, while M is the rotational part of the
local magnetic field. The magnetic field outside the metal particle is irrotational
and equals to
$ %
3 Ke − K̃m H0 · r
Hout,m = H0 + a ∇ . (304)
2Ke + K̃m r3
144 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
We assume, for simplicity, that the dielectric component of the composite is non-
magnetic, i.e., the dielectric magnetic permeability Kd = 1. Then,
3Ke
Hin,d = H0 , (305)
2Ke + 1
and
$ %
Ke − 1 H0 · r
Hout,d = H0 + a 3 ∇ . (306)
2Ke + 1 r3
As in all the EMAs described so far in this book, the effective parameters e and
Ke are determined by the self-consistent condition that the fluctuations in the fields
should vanish when averaged over all the (spherical) inclusions, i.e., Eout =
pEout,m + (1 − p)Eout,d = E0 , and Hout = pHout,m + (1 − p)Hout,d = H0 ,
where · indicates a volume averaging. Therefore, when these averagings are
carried out, they results in the following equations
e − ˜m e − d
p + (1 − p) = 0, (307)
2e + ˜m 2e + d
Ke − K̃m Ke − 1
p + (1 − p) = 0. (308)
2Ke + K̃m 2Ke + 1
These equations are completely similar to the traditional EMAs discussed in the
previous sections and Volume I. It can be seen that the skin effect results in renor-
malization of the dielectric constant and magnetic permeability of the conducting
component. Specifically, the metal dielectric constant m and magnetic permeabil-
ity Km are replaced by ˜m and K̃m given by Eqs. (293) and (301), respectively.
This fact has an important effect on the frequency dependence of the effective
parameters. For example, it is commonly accepted that the effective conductiv-
ity ge = −iωe /(4π ) of a composite is dispersion-free, when the conductivity of
metal component gm is independent of frequency and gm ω (which is typical
for the microwave and far-infrared ranges). Thus, as shown in Chapter 5 of Volume
I [see Eq. (5.62) there], the traditional EMA predicts that, ge = gm (3p − 1)/2 for
p > pc . Equation (307) yields the same result for the effective conductivity ge ,
but with the metal conductivity being renormalized according to Eq. (293), which
results in, ge = gm F (ym a)(3p − 1)/[1 − F (ym a)]. Thus, the effective conduc-
tivity has a dispersive behavior, provided that the skin effect in metal grains is
important. In the limit of very strong skin effect, δ a,√the effective conductivity
decreases with the frequency as, ge ∼ gm (δ/a) ∼ gm / ω.
Another interesting prediction is that percolation composites exhibit magnetic
properties, even if such properties are absent in each component, i.e., even if
Km = Kd = 1. In this case, the real part Ke of the effective magnetic permeability
Ke is less than one and decreases with frequency, while its imaginary part Ke has
its maximum at frequencies such that, δ ∼ a.
One can now show that the effective parameters e and Ke determine propaga-
tion of an electromagnetic wave in the metal-dielectric composites. The average
3.6. Electromagnetic Properties of Solid Composites 145
Foucault currents. Adding the electric displacement DH given by Eq. (303) to the
average displacement given by Eq. (314) yields the complete electric displacement,
4π i
Df = e E0 + ∇ × M. (316)
k
Note that the second term of Eq. (316) disappears when the skin effect vanishes,
i.e., when |ym |a → 0. We are still assuming that the linear size of the sample is
much smaller than the wavelength λ. Similarly, the average magnetic induction
Bf is given by
4π i
Bf = Ke H0 − ∇ × L. (317)
k
At this point, the Maxwell’s equations are averaged over macroscopic volume
∼ L3 , centered at point r, such that ξp L λ, yielding
∇ × E = ikBf = ikKe H0 + 4π∇ × L, (318)
∇ × H = −ikDf = −ike E0 + 4π ∇ × M. (319)
The order of the curl operation and the volume averages in Eqs. (318) and (319) can
be interchanged, as is usually done for derivation of the macroscopic Maxwell’s
equations. For example, ∇ × E = ∇ × [E(r)], where (r) indicates that the dif-
ferentiation is over the position r of the volume . Then, the Maxwell’s equations,
Eqs. (318) and (319), become
∇ × E0 (r) = ikKe H0 (r), (320)
∇ × H0 (r) = −ike E0 (r), (321)
which have the typical forms for macroscopic electromagnetism, describing
propagation of electromagnetic waves in composite media.
It is important to recognize that all quantities in Eqs. (310), (313), (314), (320),
and (321) are well-defined and do not depend on the assumptions made in the
course of their derivation. Thus, for example, M in Eq. (311) can be determined
as a magnetic moment of the Foucault currents per unit volume, so that
ik 1
M = (r × DH ) dr = (r × jH ) dr, (322)
8π 2c
where the integration now is over the volume . This definition of M is in
agreement with Eq. (302), except that it is not required that the currents IH be the
same in all the metal particles. In a similar way, one may write
ik
L = (r × BE ) dr, (323)
8π
where the integration is still over the volume , and BE = −(4π i/k)∇ × E, with
E being the local electric field. Note that L has no direct analogue in the classical
electrodynamics, since there is no such thing as loop magnetic currents in atoms
and molecules.
3.7. Beyond the Quasi-static Approximation: Generalized Ohm’s Law 147
Figure 3.13. Schematics of the model used in the computations. Electromagnetic wave
of wavelength λ is incident on a thin metal-insulator film of thickness d. The wave is
partially reflected and absorbed, and the remainder passes through the film (after Sarychev
and Shalaev, 2000).
rents flowing in between the two planes at z = −d/2 − l0 and z = d/2 + l0 are
calculated as
d/2 d/2+l0
iω −d/2
IE = − E(z) dz + m E(z) dz + E(z) dz ,
4π −d/2−l0 −d/2 d/2
(327)
iω −d/2 d/2 d/2+l0
IH = H(z) dz + Km H(z) dz + H(z) dz ,
4π −d/2−l0 −d/2 d/2
(328)
where m = 4iπgm /ω is the metal dielectric constant. We assume, for simplicity,
that the magnetic permeability Km = 1. Since the Maxwell’s equations are linear,
the local fields E(z) and H(z) are linear functions of the boundary values E1 and
3.7. Beyond the Quasi-static Approximation: Generalized Ohm’s Law 149
Maxwell’s equations that the fields and currents are connected by linear relations
given by
IE (r) = sE1 + g(n × H1 ), (338)
(Sarychev et al., 1995) [i.e., the parameter a = D/4d in Eqs. (336) and (337)]
allows one to reproduce most of the computer simulations’ results, except when a
surface polariton is excited in the corrugated film.
To simplify (341), the system of the basic equations, the electric and magnetic
fields on both sides of the film must be analyzed, namely, one must consider
these fields at a distance l0 behind the film, E1 (r) = E(r, −d/2 − l0 ), H1 (r =
H(r, −d/2 − l0 ), and at the same distance in front of the film, E2 (r) = E(r, d/2 +
l0 ), and H2 (r) = H(r, d/29 + l0 ). Then, Maxwell’s second equation, ∇ × H =
(4π/c)I, can be written as, H dl = (4π/c)(n1 · IE ), where n1 is perpendicular
to the integration contour, and the integration is over a rectangular contour which
has sides d + 2l0 and , such that the sides with length d + 2l0 are perpendicular
to the film and those with length are in the planes z = ±(d/2 + l0 ). In the limit
→ 0 this equation takes the following form
4π 4π
H2 − H1 = − (n × IE ) = − [s (n × E1 ) − gH1 ] . (343)
c c
The same procedure, when applied to Maxwell’s first equation, ∇ × H = ikH,
yields
4π 4π
E 2 − E1 = − (n × IH ) = − [m (n × H1 ) − gE1 ] . (344)
c c
Now, the electric field E1 can be expressed, using Eq. (343), in terms of the
magnetic fields H1 and H2 , while the magnetic field H1 can be expressed, using
Eq. (344), in terms of the electric fields E1 and E2 . If we substitute the resulting
expressions in the GOL, Eqs. (338) and (339), and use Eq. (333), we obtain
IE = uE, IH = wH, (345)
where E = 12 (E1 + E2 ), H = 12 (H1 + H2 ), and parameters u and w are given by
c g c g
u=− , w=− , (346)
2π m 2π s
implying that the GOL is diagonalized by introducing new fields E and H, such that
Eqs. (345) have the same form as constitutive equations of macroscopic electro-
dynamics, but with the difference that the local conductivity has been replaced by
the parameter u and the magnetic permeability K has been replaced by −4iπ w/ω.
It is then straightforward to show that, the new Ohmic
√ parameters u and w can be
expressed in terms of the local refractive index η = (r) as
c tan(Dk/4) + η tan(dkη/2)
u = −i , (347)
2π 1 − η tan(Dk/4) tan(dkη/2)
c η tan(Dk/4) + tan(dkη/2)
w=i . (348)
2π η − tan(Dk/4) tan(dkη/2)
√ √
In these equations, the refractive index η takes on values ηm = m and ηd = d
for metal and dielectric regions of the film, respectively. In the quasi-static limit,
when the optical thickness of the metal grains is small, dk|ηm | 1, but the metal
152 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
away from the film is matched with the average fields in the planes z = d/2 + l0
and z = −d/2 − l0 , i.e., E2 = e {exp[−ik(d/2 + l0 )] + r exp[ik(d/2 + l0 )]}
and E1 = e{t exp[ik(d/2 + l0 )]}. The same matching, but with the magnetic
fields, yields, H2 = (n × e) {− exp[−ik(d/2 + l0 )] + r exp[ik(d/2 + l0 )]} and
H1 = −(n × e)t exp[ik(d/2 + l0 )] in the planes z = d/2 + l0 and z = −d/2 −
l0 , respectively. Substitution of these expressions for E1 , E2 , H1 , and H2
in Eqs. (352) and (353) yields two scalar, linear equations for reflection (r) and
transmission (t) coefficients, the solution of which yields the reflectance,
. .2
. 2π .
. (ue + we ) .
. .
R ≡ |r| = .
2 . $ c % $ % . , (354)
.
. 1 + 2π u 1−
2π
we ..
. e
c c
the transmittance
. $ %2 .2
. 2π .
. 1+ .
. ue we .
. c .
T ≡ |t| = . $
2 %$ %. , (355)
. 2π 2π .
. 1+ ue 1− we ..
. c c
and the absorbance
α =1−T −R (356)
of the film. Therefore, the effective Ohmic parameters ue and we determine com-
pletely the optical properties of heterogeneous films. This analysis indicates that,
the problem of the field distribution and optical properties of the metal-dielectric
films reduces to uncoupled quasi-static conductivity problems for which extensive
theoretical analyses have already been carried out. Numerous analytical as well as
numerical methods, developed for heterogeneous media with percolation disorder
(see Chapters 4–6 of Volume I), can be employed for determining the effective
parameters ue and we of the film.
We can now consider the case of strong skin effect in the metal grains and
study the evolution of the optical properties of a semi-continuous metal film,
as the volume fraction p of the metal is increasing. When p = 0, the film is
purely dielectric and ue = ud and we = wd , where ud and wd are the dielectric
Ohmic parameters given by Eqs. (351). If we substitute ue = ud and we = wd
in Eqs. (354)–(356) and assume that the dielectric film has no losses and is op-
tically thin (i.e., dkd 1), we obtain the reflectance R = d 2 (d − 1)2 k 2 /4, the
transmittance T = 1 − d 2 (d − 1)2 k 2 /4, and the absorbance α = 0, well-known
results for a thin dielectric film (see, for example, Jackson, 1998). The losses are
also absent in the limit of full coverage, i.e., when the metal volume fraction p = 1.
Indeed, substituting the Ohmic parameters ue = um and we = wm from Eqs. (350)
in Eqs. (354)–(356) yields, R = 1, T = 0, and α = 0. Note that in the limits p = 0
and p = 1, the optical properties of the film do not depend on the particle size D,
because properties of the dielectric and continuous metal films should not depend
on the shape of the metal grains.
154 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
1
T (pc ) = " 2 , (359)
1 + d
"
2 d
α(pc ) = " 2 , (360)
1 + d
where, d = 1 + 2d d/D, as before. When metal grains are oblate enough that
d d/D 1 and d → 1, one obtains the universal result
R = T = 1/4, α = 1/2, (361)
implying that there is effective absorption in semi-continuous metal films even for
the case when neither dielectric nor metal grains absorb light energy. The effective
absorption in a loss-free film means that the electromagnetic energy is stored in
the system, and that the amplitudes of the local electromagnetic field can diverge.
In practice, due to non-zero losses, the local field saturates in any semi-continuous
metal film.
To determine the optical properties of semi-continuous films for arbitrary metal
volume fraction p, the EMA can be used which then yields the following equations,
u2e − pue (um − ud ) − ud um = 0, (362)
we2 − pwe (wm − wd ) − wd wm = 0, (363)
where p = (p − pc )/pc (pc = 1/2). Equation (363) indicates that, when the
skin effect is strong and wm and wd are given by Eqs. (350) and (351), then
|we | c for all metal volume fractions p, and therefore we can neglect we in
Eqs. (354) and (355). Moreover, compared with um , ud can also be neglected in the
second term of Eq. (362). Thus, introducing the dimensionless Ohmic parameter
ue = (2π/c)ue allows us to rewrite Eq. (362) as
λp
ue − 2i u − d = 0.
2
(364)
πD e
3.8. Piecewise Linear Transport Processes 155
"
At p = pc = 1/2 (i.e., where p = 0), Eq. (364) predicts that, ue (pc ) = d ,
which coincides with the exact result, Eq. (357), and those given by Eqs. (358)–
(360). For p = pc , Eq. (364) predicts that
#
$ %
λp λp 2
ue = i + d − , (365)
πD πD
"
which becomes purely imaginary for |p| > πD d /λ. Then, α = 1 − |ue |2 /
|1 + ue |2 − 1/|1 + ue |2 = 0 "
(recall that we was neglected). In the vicinity of pc ,
namely, for |p| < (π D/λ) d , the effective Ohmic parameter ue has a non-
vanishing real part, and therefore
"
2 d − [λp/(π D)]2
α= " , (366)
1 + d + 2 d − [λp/(π D)]2
which has a well-defined maximum at pc , with the width of the maximum being
inversely proportional to the wavelength. These predictions were confirmed by
extensive numerical simulations. They are also in agreement with the experimen-
tal data (see Sarychev and Shalaev, 2000, for detailed discussions). Note that the
parameters ue and we can be determined experimentally by measuring the am-
plitudes and phases of the transmitted and reflected waves using, for example, a
waveguide technique (see, for example, Golosovsky et al., 1993 and references
therein), or by measuring the film reflectance as a function of the fields E1 and H1 .
which depends on the size of the sample. Bingham fluids are viscous and behave
like Newtonian fluids if the shear stress applied to them is larger than a critical
value, but do not flow if the stress is less than the threshold value. An example
of such fluids, already described in Section 9.3 of Volume I, is foam. In order to
mobilize foam and force it to flow, the applied pressure must exceed a critical
value; otherwise it will not flow.
Let us consider a 2D or 3D resistor network in which every bond is characterized
by the following current-voltage relation,
g(v − vc )n , v > vc ,
i= (367)
0 v ≤ vc ,
where vc is the critical voltage or threshold for the onset of transport. As in the
case of strong and weak nonlinearities, we take g to be a generalized bond con-
ductance which, in general, can vary from bond to bond. On the other hand, in any
physical situation involving a disordered material, one expects a distribution of
the thresholds vc , because due to a variety of factors, different parts of a material
may become conductive beyond different thresholds. Therefore, one may make
the simplification that, instead of assuming g to be a statistically-distributed vari-
able, vc is assumed to be randomly distributed which, for the sake of simplicity, is
assumed to be distributed uniformly in (0, 1). The conductivity g is then the same
for all bonds, and therefore its numerical value is irrelevant (we assume g = 1).
The questions that we ask are:
(1) What is the critical voltage Vc in order to have macroscopic transport in the
network, and
(2) how do the macroscopic current I and the effective conductivity ge of the
network vary with the applied voltage? The piecewise linear process that we
study here is reversible, i.e., if I is lowered the conducting bonds become
3.8. Piecewise Linear Transport Processes 157
where vci is the critical voltage of bond i, and the sum is taken over all paths
between the two terminals of the network. Equation (368) immediately necessitates
the concept of an optimal path between the two terminals of the network (see, for
example, Cieplak et al., 1994,1996; Porto et al., 1997). Obviously, if the applied
voltage is larger than a final or the last voltage threshold Vl , all bonds of the network
will be conducting, and one will have the usual linear transport in which the
current I is simply proportional to V . Therefore, one generally has three regimes
of interest:
(1) If V < Vc , then enough bonds have not become conducting to form a sample-
spanning cluster, and therefore no macroscopic transport takes place. Hence,
I = 0 and ge =0.
(2) If Vc < V < Vl , then enough bonds have become conducting that make macro-
scopic transport possible, while some of the bonds are still not conducting. We
expect I to depend nonlinearly on V − Vc , because this is precisely the regime
in which the effect of nonlinearity (random voltage thresholds) should manifest
itself. As we show below, this is indeed the case (note that in linear transport
above pc , I always varies linearly with V ).
(3) If V > Vl , then every bond of the network is conducting, the normalized
effective conductivity is ge = 1, and I depends linearly on V again.
Figure 3.15. Effective conductivity of the square network with piecewise linear resistors
with a threshold, versus the applied voltage (after Sahimi, 1993a).
160 3. Nonlinear Conductivity, Dielectric Constant, and Optical Properties
Figure 3.16. The I − V characteristics of the square network of Figure 3.15 (after Sahimi,
1993a).
EMA. While computer simulations indicate that, Vc 0.29, the EMApredicts that,
Vc = 3/8 = 0.375. Roux et al. (1987) used a transfer-matrix method described
in Section 5.14.2 of Volume I and estimated that for a square network, tilted at
45◦ , one
√ has, Vc 0.23 [in general, for the square network, Vc (tilted)=Vc (non-
tilted)/ 2]. This difference can be explained by the fact that, because the resistor
network that Roux et al. (1987) used in their simulations was tilted, their network
is different from a non-tilted one, since the distribution of currents in the bonds
of their network is isotropic, whereas the same distribution is anisotropic in a
non-tilted network. The difference is due to the fact that the bonds of a non-
tilted network that are perpendicular to the direction of the applied voltage receive
much less current than those that are aligned with it. As a result, formation of a
sample-spanning conducting cluster is easier in a tilted network than in a non-
tilted one, implying that the critical voltage Vc for a tilted network should be
smaller than that of a non-tilted one. Thus, such local anisotropies, which usually
have no consequence for macroscopic properties of linear transport processes,
are important in a nonlinear system, such as what is described here. Moreover,
according to Eq. (375), in the nonlinear regime, the macroscopic current I varies
quadratically with V − Vc , which is in agreement with the simulations of Roux
and Herrmann (1987), Eq. (369).
3.8. Piecewise Linear Transport Processes 161
Summary
Using the discrete models, we described and analyzed several types of nonlinear
transport and optical properties of disordered materials. As our analyses indicate,
the interplay of nonlinearity and the disordered morphology of a material gives
rise to a rich set of phenomena that are absent in linear transport processes in
the same material. In particular, strong heterogeneity, such as percolation-type
disorder, enhances the nonlinear response of a material, and shrinks the range of
the parameter space in which the material behaves linearly, and hence opens up
the possibility of developing composite materials with highly unusual and useful
properties.
4
Nonlinear Rigidity and Elastic Moduli:
The Continuum Approach
4.0 Introduction
In this chapter we consider nonlinear mechanical properties of heterogeneous
materials. This class of problems has many applications that will be described
throughout this chapter. However, to give the reader an interesting and somewhat
unusual application of this class of phenomena, we consider the following prob-
lem. It has been observed (Gordon, 1978) that extensible biological tissues, such
as skin, are difficult to tear, even though their specific work of fracture (see the
discussions in Chapters 6 and 7) is not large compared to those of materials that
tear easily. For example, the fracture toughness of animal membranes is around
1-10 kJm−2 , an order of magnitude smaller than aluminum foil which tears easily.
Gordon reasoned that this difference is due to the markedly different shape of the
stress-strain diagram of such materials. Figure 4.1 presents schematic stress-strain
curves for extensible biological tissues, rubber, and the standard Hookean solid for
which the diagram is a straight line. The small-strain portion of the J-shaped curve
of the biological material is indicative of lack of shear connection in the material,
i.e., absence of shear stiffness in what are anisotropic solids. This diagram provides
an explanation as to why such materials are difficult to tear, because it is difficult
to concentrate energy into the path of a putative crack. Note also the difference be-
tween the stress-strain diagrams for rubbers and the biological materials: For small
strains, the rubber’s curve is not J-shaped, which may also explain why we cannot
replace human body arteries or veins by rubber tubes. We also remind the reader
that when Nature does want fracture and tear to happen, as in, for example, amniotic
membranes and eggshells, the stress-strain diagrams are Hookean linear elastic!
Studies of heterogeneous materials with nonlinear constitutive behavior go back
to at least Taylor (1938) who studied the plasticity of polycrystals, and to the
subsequent work by Bishop and Hill (1951a,b) and Drucker (1959) who investi-
gated the behavior of ideally plastic polycrystals and composite materials. Over
the past decade or so, numerical simulations of nonlinear materials with periodic
microstructures have been carried out (see, for example, Christman et al., 1989;
Tvergaard, 1990; Bao et al., 1991), as well as materials with more general mi-
crostructures (see, for example, Brokenborough et al., 1991; Moulinec and Suquet,
1995). Such efforts will be briefly described in this chapter where we make com-
parison between the theoretical predictions and the numerical simulation results.
4.1. Constitutive Relations and Potentials 163
5JHAII
Strain
The main advantage of such simulations is that they provide accurate description
of the system under study, and yield useful insight into their properties. Their main
disadvantage is that they require very intensive computations, especially when the
material’s microstructure is disordered.
In this chapter we describe and discuss recent advances in understanding
the effective mechanical properties of disordered materials with constitutive
nonlinearity. Although one may argue that numerical techniques, such as the
finite-element methods, represent some form of discrete approach to this class
of problems, to our knowledge very little work has been done using the discrete
network models of the type that we have so far described and discussed for esti-
mating various transport properties of disordered materials. Therefore, the main
focus of this chapter is on the theoretical developments based on nonlinear contin-
uum models of disordered materials. These theoretical approaches represent the
mechanical analogues of those described in Chapter 2 for estimating the effective
conductivity and dielectric constant of nonlinear materials. Thus, the methods that
we describe in this chapter are based on rigorous variational principles which, in
addition to possessing mathematical rigor, have the advantage of leading to bounds
and relatively accurate estimates for the mechanical properties. As described and
discussed in Chapter 2, such variational principles allow one to obtain estimates of
the effective energy densities of nonlinear materials in terms of the corresponding
information for linear composites with the same microstructure. A large portion
of our analyses and discussions in this chapter is based on an excellent review by
Ponte Castañeda and Suquet (1998).
with ni ≥ 1 and τi0 being the creep exponent and reference stress of the ith slip
system, respectively, and γ 0 is a reference strain rate. The constitutive relations
166 4. Nonlinear Rigidity and Elastic Moduli: The Continuum Approach
(14) and (15) can then be expressed in terms of the convex potential for the crystal:
M
u(c) (σ ) = ψi (σ : µi ), (17)
i=1
such that
∂u(c)
= . (18)
∂σ
The limit ni → ∞ corresponds to a rigid, ideally plastic crystal, with a strength
domain given by
P = {σ , τi ≤ τi0 , i = 1, . . . , M}. (19)
We can assume, more generally, that the potential w can be expressed by
w() = F (E), (20)
where F is an appropriately-selected function, and E is a fourth-rank tensor which
is defined by
1
E= ⊗ , (21)
2
and possesses the usual diagonal symmetry and positive-definitiveness property
of an elasticity tensor. The function F is then defined on the space of fourth-rank
tensors P that have diagonal symmetry, so that the constitutive relation (1) can be
written as
σ = Ls (E) : , (22)
with
∂F
Ls (E) = , (23)
∂P
being the secant modulus tensor of the material, which also has diagonal symmetry.
Given Eq. (20), the dual potential u can be expressed as
1
u(σ ) = G(S), S= σ ⊗ σ, (24)
2
where G is a function of fourth-rank tensors S. In terms of the secant compli-
ance tensor of the material, the constitutive relation (3) may be expressed in the
following form
∂G
= Ms (S) : σ , Ms (S) = . (25)
∂S
As an example, consider crystalline materials. First, note that
τi2 = 2Mi :: S, Mi = µi ⊗ µi , (26)
4.2. Formulation of the Problem 167
so that
M
√
u(σ ) = gi (2Mi :: S) = G(S), gi (x) = ψi ( x), (27)
i=1
and the compliance tensor is given by
M
Ms (S) = 2 αi Mi , αi = gi (τi2 ). (28)
i=1
N
= σ = φi σ i , (38)
i=1
where
S1 (E) = {v = E · x on ∂}, (41)
while the second one is in terms of the minimum complementary energy, according
to which τ is the solution of the problem
inf u(σ ), (42)
τ ∈S2 ( )
with
S2 () = {τ , ∇ · τ = 0, in , τ = }. (43)
4.3. The Classical Variational Principles 169
Then, since the infimum problem in (40) defines the average strain energy in the
material, the effective strain-energy potential He is defined as
He (E) = inf w[(v)], (44)
v∈S1 (E)
so that
5 $ %6 5 $ %6
∂He ∂w ∂u ∂u
= [(u)] : = σ : . (45)
∂E ∂ ∂E ∂E
However, since ∂u/∂E = U · x, where U is the identity tensor in the space of
fourth-rank tensors, it follows from Eq. (39) that
∂He
= σ : U = , (46)
∂E
which defines the effective stress-strain relation for the material. Similarly, the
effective stress-energy potential He∗ is defined as
He∗ () = inf u(τ ), (47)
τ ∈S1 ( )
in terms of which,
∂He∗
E= . (48)
∂
Both He and He∗ are convex functions. Furthermore, it can be shown (Suquet,
1987; Willis, 1989a) that they are in fact the (Legendre) dual functions, such that
He (E) + He∗ () = w() + u(σ ) = σ : = : E, (49)
and that they correspond to the boundary conditions (35). Adopting the boundary
condition (36) would lead to different pairs of dual potentials. However, under
the assumption that the potentials w and u are strictly convex, the two types of
boundary conditions are equivalent for the representative volume element, and
are also equivalent to the periodic boundary conditions used in the theory of
homogenization (see, for example, Sanchez-Palencia, 1980).
As an example, consider the limiting case of rigid, ideally plastic materials for
which the potentials are convex, but not strictly. In this limit, which requires special
treatment (Bouchitte and Suquet, 1991), He is a positively-homogeneous function
of order one in E, usually referred to as the plastic dissipation function. It may
also be useful to introduce the effective strength domain of the material, defined
as (Suquet, 1983)
Pe = { such that there exists σ (x) with σ = and
and that
0 if ∈ Pe
He∗ () =
+∞ otherwise
and
N
He∗ () ≤ u() = φi wi (), (52)
i=1
or, equivalently,
N ∗ N
φi ui (E) ≤ He (E) ≤ φi wi (E), (53)
i=1 i=1
where superscript ∗ denotes the convex dual function. In the context of polycrys-
talline materials, the bounds (51) and (52) are commonly referred to as the Taylor
(1938) and Sachs (1928) bounds, respectively. For example, the Reuss and Voigt
bounds for incompressible, isotropic power-law phases are given by
$ % $ %
(σ 0 )−n −m 0 Eeq m+1 σ 0 0 Eeq m+1
≤ H e (E) ≤ . (54)
m+1 0 m+1 0
Since the Voigt and Reuss bounds incorporate only limited information on the
morphology of a material—the volume fractions of the phases—they are not very
useful, particularly when the contrast between the phases is large. In fact, they can
be shown to be exact only to first order in the contrast between the properties of
the phases.
(1985), following the earlier work of Willis (1983), which we now describe and
discuss.
Let w 0 be the potential function of a linear, homogeneous reference material
with uniform modulus tensor L0 , such that
1
w0 () = : L0 : , (55)
2
and assume that the difference potential (w − w0 ) is a concave function, so that
the concave polar of this difference is defined as (see Ponte Castañeda and Suquet,
1998)
7 8
(w − w0 )∗ (x, τ ) = inf τ : − w(x, ) − w0 () .
The concavity of (w − w 0 ) results in
7 8
w(x, ) − w 0 () = inf τ : − (w − w 0 )∗ (x, τ ) , (56)
τ
Substituting (56) for w in Eq. (47) and interchanging the order of the infima over
and τ , one arrives at
7 8
He (E) = inf inf w 0 [(v)] + τ : (v) − (w − w 0 )∗ (x, τ ) . (57)
τ v∈S1 (E)
It then follows that minimizing the displacement field u is equivalent to finding
the solution to the following boundary value problem:
∇ · L0 : (u) = −∇ · τ , u ∈ S1 (E). (58)
If one utilizes the Green function G0 associated with the system (58) in the domain
, one obtains the following expressions for the strain tensor,
= E − 0 ∗ τ , (59)
where, as before, E is the average strain over , and
2 3
∗τ =
0
0 (x, x ) : τ (x ) − τ dx , (60)
with
∂ 2 G0ik
ij0 kl = .
∂xj ∂xl
(ij ),(kl)
is usually used. Since φi = mi (x) denotes the volume fraction of the phase
i, and given the fact that the average of a tensor T over phase i is given by,
Ti = (mi /φi )T, it follows from (59) and (60) that
1
N
i = E − ij : τ j , (62)
φi
j =1
where
5 6
2 3
ij = mi (x) mj (x ) − φj 0 (x, x )dx , i, j = 1, · · · , N (63)
are tensors that depend only on the microstructure of the material and L0 , and ij
are symmetric (Kohn and Milton, 1986) in i and j and are not all independent,
since they satisfy the relations
N
N
ij = ij = 0.
i=1 j =1
1
N
∂
(wi − w )∗ (τ i ) +
0
ij : τ j = E, i = 1, . . . , N, (65)
∂τ i φi
j =1
so that, from Eqs. (46), (64), and (65) one finally obtains [by replacing the
inequality in (64) by an equality] an approximate stress-strain relation:
= 0 : E + τ , (66)
where the τ i are obtained from Eqs. (65).
The upper bound (64) for He (E), which was first given by Ponte Castañeda
and Willis (1988), can be written in an alternative form (Willis, 1991) by noting,
4.4. Variational Principles Based on a Linear Comparison Material 173
through the use of (62), that the optimality conditions (65) can be rewritten in the
form
∂
i = wi − w 0 (τ i )
∂τ i ∗
1
N
∂
i + ij : (wj − w 0 )∗ () = E, (68)
φi ∂j
j =1
N
He (E) ≥ w0 (E) + φi [2τ i : (E − i ) + (wi − w 0 )∗∗ (i )], (69)
i=1
(1991a) for materials with isotropic phases and by Suquet (1993a) for compos-
ites with power-law phases. Moreover, a hybrid of the Talbot–Willis and Ponte
Castañeda procedures, using a linear thermoelastic comparison material, was pro-
posed by Talbot and Willis (1992). These procedures can in fact be shown to be
equivalent under appropriate hypotheses on the local potentials. An important ad-
vantage of the variational procedures that involve linear comparison materials is
that, they can not only produce the nonlinear Hashin–Shtrikman-type bounds of the
Talbot–Willis procedure directly from the corresponding linear Hashin–Shtrikman
bounds, but also yield higher-order nonlinear bounds, such as Beran-type bounds,
as well as other types of estimates. The application of this technique to deriving
bounds and estimates for the effective nonlinear conductivity and dielectric con-
stant of materials was described and discussed in Chapter 2. We now describe the
analogous results for the effective nonlinear mechanical properties of materials.
with the functions fi , characterizing the deviatoric behavior of the material (see
Chapter 7 of Volume I), being defined by the relations, fi (p) = ϕi (eq ) for p =
2 . The functions f are assumed to be concave functions of p, such that f (p) =
eq i i
−∞ for p < 0, fi (0) = 0, and fi → ∞ as p → ∞. By definition, the concave
dual function of fi is given by
fi∗ (q) = inf {pq − fi (p)} = inf {pq − fi (p)} .
p p>0
Then, using (71) with q = 3µ0 /2, one finds that the potential of the nonlinear
material w is given by the exact equation,
7 8
w(x, ) = inf w 0 (x, ) + v(x, µ0 ) , (73)
µ0 (x)>0
where
N
3
v(x, µ0 ) = mi (x)vi [µ0 (x)], with vi (µ0 ) = −fi∗ ( µ0 ), (74)
2
i=1
from which one obtains, by interchanging the order of the infima over and µ0 ,
7 8
He (E) = inf He0 (E) + V (µ0 ) , (77)
µ0 (x)
where V (µ0 ) = v[x, µ0 (x)], and He0 is the effective potential of the linear
comparison material (Ponte Castañeda, 1992a):
where
One can then show that (Ponte Castañeda and Suquet, 1998)
He (E) =
⎧ ,$ -(1−m)/2 ⎫
⎨ %(m+1)/(m−1) ⎬
1 1 3
inf H 0
(E) (m+1)/2
µ 0
(σ 0 2/(1−m)
) ,
m + 1 ( 0 )m µ0 (x)>0 ⎩ e
2 ⎭
(81)
and that
⎧ , -(1−n)/2 ⎫
0 ⎨ (σ 0 )2n/(n−1) ⎬
He∗ () = sup (He∗ )0 ()(n+1)/2 .
n+1 µ0 (x)>0 ⎩ 6(µ0 )(n+1)/(n−1) ⎭
(82)
where He0 is the effective potential of the linear comparison material defined by
the local potential (85), and V (L0 ) = v[x, L0 (x)], given by
Equation (86) expresses the nonlinear effective properties of the material in terms
of two functions which are, (1) He0 , the elastic energy of a fictitious linear het-
erogeneous solid, called the linear comparison material, that consists of phases
with stiffness L0 (x), and (2) v(x, ·), the role of which is to measure the difference
between the non-quadratic potential w(x, ·) and the quadratic energy of the linear
comparison solid. The linear comparison solid is selected from amongst all the
possible comparison materials by solving the optimization problem (86).
Equation (86), which is exact, is strictly equivalent to the variational represen-
tation of He given by (44). However, determining the exact solution of (86) is not
possible. The difficulty lies in the precise determination of the energy He0 for a
linear comparison solid consisting of infinitely many different phases, about which
very little is known. For this reason, except for very simple microstructures, such
as laminates considered in Section 2.3, the optimal solution of (86) is not known,
and only sub-optimal solutions can be determined. We now consider application
of these principles to a few classes of materials.
M
(k)
Gi (S) = g(k) 2Mi :: S ,
k=1
where, as before
(k) (k) (k)
Mi = µi ⊗ µi .
178 4. Nonlinear Rigidity and Elastic Moduli: The Continuum Approach
with
N (k) (k) (k)
(k) 2 k=1 αi Mi , αi > 0,
Mi = (88)
+∞, otherwise.
Therefore, the corresponding local stress potential for the linear comparison
polycrystalline material is given by
M . . . .
1 (k) . (k) .2 . (k) .2
u0i (σ ) = σ : Mi : σ = αi .τi . , .τi . = 2Mi :: σ . (89)
2
k=1
which was first derived by deBotton and Ponte Castañeda (1995). The functions
(k)
αi (x), k = 1, . . . , M, are defined over the region in space that is occupied by
the crystals with fixed orientation i.
where Ĥe0 is the same as He0 in (86), but with L0 replaced by L̂0 . Evaluating the
minimum over t yields an exact representation for He :
(1−m)/2
2 (m+1)/2 1 + m
He (E) = inf He (E)
0 0
V (L ) , (91)
m + 1 L̂0 >0 1−m
where the hat notation has been deleted for simplicity. The analogous representa-
tion for He∗ is given by
(1−n)/2
2 (n+1)/2 n + 1
He∗ () = ∗ 0
sup (He ) () 0
V (L ) . (92)
n + 1 L0 >0 n−1
and
5 $ %6
3 0
V (L0 ) = − f ∗ µ
2
2/(1−m) ,$ % -
1−m 1 3 0 (m+1)/(m−1) 0 2/(1−m)
= µ (σ ) .
1 + m 2( 0 )m 2
He∗ () =
⎧ N M , ⎫
⎨ 2n/(n−1) - (1−n)/2 ⎬
γ0 (k)
sup (H∗ )0 ()(n+1)/2
(k)
φi [αi ](n+1)/(n−1) τ 0 ,
n + 1 α (k) >0 ⎩ e i ⎭
i i=1 k=1 i
(94)
for i = 1, · · · , N and k = 1, · · · , M.
180 4. Nonlinear Rigidity and Elastic Moduli: The Continuum Approach
where the tensors L0i are assumed constant. In this manner an upper bound for He ,
given by
N
He (E) ≤ inf He0 (E) + φi vi (L0i ) , (98)
L0i >0, i=1,···,n i=1
with (He∗ )0 now being the effective stress potential associated with the same linear
comparison material as for He0 given above, i.e., one with the same microstructure
as the nonlinear material, but with the domains i occupied by linear phases with
compliances M0i . From Eq. (101) one obtains
∂He∗
E= () = M0e (M0i ) : . (102)
∂
It also follows from (100) that
N
N
∗ ∗
He () ≥ inf φi Gi (Si ) = φi Gi (S)i , (103)
τ ∈S2 ( ) i=1 i=1
where Si∗ = Si = 12 σ ⊗ σ i is the second moment of the stress field in phase i
of the linear comparison material. The compliances M0i of the comparison mate-
rial are determined as the solution of the optimization problem (100), which can
alternatively be written in terms of the solution of the following nonlinear problem
for the variables Si∗ :
∂Gi ∗ 1 ∂M0e
M0i = (S ), Si∗ = : : . (104)
∂S i 2φi ∂M0i
where the functions vi are defined by Eq. (75). The upper bound (105), as well as
the analogous lower bound,
1
N
∗
He () ≥ sup : Me (µi ) : −
0 0 0
φi vi (µi ) , (106)
µ0 >0, i=1,···,N 2
i i=1
were first derived by Ponte Castañeda (1991a). Based on the associated optimality
conditions (deBotton and Ponte Castañeda, 1992,1993), it can be shown that
9
N N
(m) 2 eq 2
He ≤ φ i K i i + φi ϕi i , (107)
2
i=1 i=1
φi (m) 2
N N
1 eq 2
He∗ ≥ σ + φi ψi σi , (108)
2 Ki i
i=1 i=1
182 4. Nonlinear Rigidity and Elastic Moduli: The Continuum Approach
where
1/2 1/2
(m) 1 ∂L0e (m) 1 ∂M0e
i = E: :E , σi = : : , (109)
9φi ∂Ki0 φi ∂(1/Ki0 )
and
1/2 1/2
eq 1 ∂L0e eq 3 ∂M0e
i = E: :E , σi = : : . (110)
3φi ∂µ0i φi ∂(1/µ0i )
These simplified bounds were first given by Suquet (1995,1997).
For power-law materials, one obtains,
⎧ N (1−m)/2 ⎫
1 1 ⎨ $ 3 %(m+1)/(m−1) ⎬
He (E) ≤ inf H 0
(E)(m+1)/2
φ i µ0
(σ 0 2/(1−m)
) ,
m + 1 ( 0 )m µ0i >0 ⎩ e
2 i i ⎭
i=1
(111)
and
⎧ N (1−n)/2 ⎫
0 ⎨ ⎬
He∗ () ≥ ∗ 0
sup (He ) ()(n+1)/2 0 (n+1)/(1−n) 0 2n/(n−1)
φi (6µi ) (σi ) ,
n + 1 µ0 >0 ⎩ ⎭
i i=1
(112)
which were also derived by Suquet (1993a).
These nonlinear equations can be expressed more explicitly in terms of the slip
(k)
compliances αi and the corresponding second moment of the resolved shears,
1/2
(k) (k)
τ̄i = 2Mi :: σ̄i .
functions of t. Since t is small, one can write down a perturbation series expansion
of He about t = 0, given formally by
∂He 1 ∂ 2 He
He (E, t) = He (E, 0) + t (E, 0) + t 2 (E, 0) + O(t 3 ). (120)
∂t 2 ∂t 2
The problem to be solved for ut is given by
∂w
∇· [x, (ut ), t] = 0, ut ∈ S1 (E). (121)
∂
If we differentiate (121), we find that u̇t = ∂ut /∂t is the solution of the following
system of equations
∇ · [Lt : (u̇t )] + ∇ · τ t = 0, u̇t ∈ S1 (0). (122)
where
∂ 2w ∂ 2w ∂
Lt = [x, (ut ), t], τ t = [x, (ut ), t] = (δw)[x, (ut ), t].
∂∂ ∂t∂ ∂
Therefore,
5 6 5 6
∂He ∂w ∂w
(E, t) = [x, (ut ), t] : (u̇t ) + [x, (ut ), t] . (123)
∂t ∂ ∂t
The first term of (123) vanishes due to Eq. (39) (the so-called Hill’s lemma), and
therefore,
∂He
(E, t) = δw[x, (ut )]. (124)
∂t
Using Eq. (121), one obtains
5 6
∂ 2 He ∂
(E, t) = (δw)[x, (ut )] : (ut ) = − (u̇t ) : Lt : (u̇t ) . (125)
∂t 2 ∂
Because the material is homogeneous for t = 0, u0 = E · x, and therefore
He (E, 0) = w 0 (E),
∂He
(E, 0) = δw(E), (126)
∂t
∂ 2 He
(E, 0) = −(u̇0 ) : L0 : (u̇0 ),
∂t 2
where
∂ 2 w0
L0 = (E). (127)
∂∂
Here, u̇0 is the solution of the linear elasticity problem,
∇ · [L0 : (u̇0 )] + ∇ · τ = 0, u̇0 ∈ S1 (0), (128)
with
∂
τ (x) = (δw)(x, E).
∂
186 4. Nonlinear Rigidity and Elastic Moduli: The Continuum Approach
Since the modulus tensor L0 is constant, the problem posed by (128) is a linear
elasticity problem for a homogeneous material with a distribution of body forces
determined by τ .
If the material consists of N homogeneous phases, then τ is piecewise constant,
i.e., it is a constant in each phase, with
N
∂
τ (x) = mi (x) τ i , τi = (δwi )(E),
∂
i=1
2
(E, 0) = − τ i : ij : τ j , (129)
∂t
i=1 j =1
where the microstructural tensors ij are defined by Eq. (63). Therefore (Suquet
and Ponte Castañeda, 1993),
1 2
N N
He (E, t) = w(E) − t τ i : ij : τ j + O(t 3 ). (130)
2
i=1 j =1
Similar to the case of the effective conductivity and dielectric constant of nonlin-
ear heterogeneous materials that was discussed in Chapter 2, this method is based
on a Taylor expansion for the phase potentials wi . Thus, introducing reference
strains E(i) , the Taylor expansion for wi about E(i) is given by
1
wi () = wi [E(i) ] + ρ i : [ − E(i) ] + [ − E(i) ] : Li : [ − E(i) ], (132)
2
where ρ i and Li are, respectively, an internal stress and a tangent modulus tensor,
with components
∂wk (k) ∂ 2 wm
(ρk )ij = [E ], (Lm )ij kl = [Ẽ(m) ]. (133)
ij ∂ij ∂kl
Li depends on the strain Ẽ(i) = λ(i) E(i) + [1 − λ(i) ], where λ(i) depends on
and is such that 0 < λ(i) < 1.
In terms of the average E and fluctuating components of = E + , Eq. (132)
is rewritten as
1
wi (E + ) = νi + τ i : + : Li : , (134)
2
where
1
νi = wi [E(i) ] + ρ i : [E − E(i) ] + [E − E(i) ] : Li : [E − E(i) ], (135)
2
τ i = ρ i + Li : [E − E(i) ]. (136)
Making the approximation that the strains Ẽ(i)
are constant in each phase, the
effective potential He of the nonlinear material is then estimated as
N
He (E) H̃e (E) = φ i νi + P , (137)
i=1
where
5 6
1
P =
inf (v ) : L : (v ) + τ : (v ) . (138)
v ∈S1 (0) 2
τ and L(x) are defined by equations similar to (61). The advantage of approxi-
mation (137), relative to the exact result (44), is that it requires only the solution
of a linear problem for an N -phase thermoelastic material, as defined by the
Euler–Lagrange equations of the variational problem P in (138):
∇ · [L : (u )] = −∇ · τ , u ∈ S1 (0). (139)
Estimates for N-phase linear-thermoelastic materials can, in general, be obtained
by appropriate extension of the corresponding methods for N -phase linear-elastic
composites (see, for example, Willis, 1981). Similar, but not equivalent, repre-
sentations for the effective mechanical properties of nonlinear materials, which
also utilize heterogeneous thermoelastic reference materials, were proposed by
Molinari et al. (1987) and Talbot and Willis (1992).
188 4. Nonlinear Rigidity and Elastic Moduli: The Continuum Approach
Equation (137) provides an estimate for He for any choice of E(i) and Ẽ(i) , if
we supply it with an estimate for P . A plausible approximation for the E(i) is to set
them equal to the averages of the strain field over the phases i. However, because
the exact strain field is not known, the approximate field , as determined by (139),
is used, so that
E(i) = i , (140)
where, i = E + i . Equation (140) is a reasonable choice because the strain
in phase i is expected to fluctuate about its average in phase i in such a way that
large deviations would only be expected in regions of relatively small measure.
The following identity, obtained from Eq. (139),
1 ∂P
i = , (141)
φi ∂τ i
which can be used to obtain i directly from P , via
1 ∂P
i = E + , (142)
φi ∂τ i
is also useful, since the reference strains E(i) may also be computed from P by
means of Eqs. (140) and (142). It can also be shown that Eq. (140) provides the
optimal choice for E(i) in the sense that, estimate (137) for He is stationary with
respect to the E(i) .
One important consequence of stationarity of Eq. (140) is that the overall stress-
strain relation (46) for the material may be approximated as
N
1 ∂ Ẽ(i)
= φi ρ i + ( − i ) : Ni : ( − i )i : , (143)
2 ∂E
i=1
the second-order estimates based on the estimate He are more accurate than the
analogous estimates for He∗ .
where L̂ and M̂ are two microstructural tensors that are independent of µ01 , then,
the estimate (152) for He and the corresponding estimate for He∗ reduce to
eq eq
He (E) ≤ φ1 ϕ1 1 , He∗ () ≥ φ1 ψ1 σ1 , (155)
4.7. Applications of Second-Order Exact Results 191
where
# #
eq 1 eq 3
1 = E : L̂ : E, σ1 = : M̂ : . (156)
3φ1 φ1
Use of any lower bound or any estimate for the effective compliance tensor of a
rigidly-reinforced material with an isotropic matrix then leads to corresponding
lower bounds and estimates for the effective stress potential of the corresponding
nonlinear, rigidly-reinforced composites. When the matrix phase is also incom-
pressible (i.e., when K1 → ∞), the resulting material is also incompressible and
the corresponding estimates for He and He∗ can be written in a form similar to
(155) and (156), with Em = 0, in terms of appropriate microstructural tensors L̂
and M̂ = (L̂)−1 . For example, the Hashin–Shtrikman estimates can be interpreted
as appropriate variational estimates for particulate microstructures, and thus the
corresponding nonlinear results can be thought of as appropriate variational esti-
mates for particulate microstructures. Thus, (155) and (156), with the inequality
replaced by an approximate equality, yield estimates for He and/or He∗ . In particu-
lar, for spherical particles that are distributed with statistically-isotropic symmetry,
the following estimate should be used:
&
eq 1 3
1 = 1 + φ2 Eeq , (170)
φ1 2
This “lower estimate” was proposed by Ponte Castañeda (1991b, 1992a) for
isotropic, rigidly-reinforced composites and generalized by Talbot and Willis
(1992) for anisotropic materials. Talbot and Willis (1992) and Li et al. (1993)
also presented predictions for aligned spheroidal inclusions. Gărăjeu and Suquet
(1997) also discussed an application to rigidly-reinforced materials.
with
&
σ10 2 1
≥ 1 + dφ2 ,
0
σ2 d +2 2
and
&
σe0 1
= 1 + dφ2 , (182)
σ10 2
where
&
σ10 2 1
≤ 1 + dφ2 .
0
σ2 d + 2 2
These results, which may be interpreted as approximate estimates for materials
with particulate microstructures, are upper bounds for composites with morpholo-
gies for which the Hashin–Shtrikman lower bound for µe is exact (for example,
sequentially-laminated composites; see Chapter 2). Note that the estimate (182)
predicts that the strengthening effect of the inclusions (when they are stronger than
the matrix) saturates after a certain finite increase in the strength of the inclusions.
This is a consequence of the non-hardening character of the matrix phase, which
would be expected to carry all the deformation, for sufficiently strong (but still
non-rigid) inclusions.
198 4. Nonlinear Rigidity and Elastic Moduli: The Continuum Approach
where y = µ01 /µ02 , and the tensor M̂0e = µ01 M0e is the (normalized) effective com-
pliance of the fiber-reinforced linear comparison material with incompressible and
isotropic phases. In general, this result requires numerical computation, but for
transverse and longitudinal shear, the result simplifies to expressions similar in
form to (180)–(182) with d = 2. Similarly, for (axisymmetric) uniaxial tension,
one obtains
σe0 = φ1 σ10 + φ2 σ20 , (184)
in agreement with the Voigt estimate. These results are due to Ponte Castañeda
and deBotton (1992) and Moulinec and Suquet (1995); see also deBotton (1995).
In a similar way, one may obtain second-order estimates for the effective me-
chanical properties of this class of nonlinear materials. For example, for two-phase,
rigid, perfectly plastic materials with statistically-isotropic microstructures [or
with isotropic distributions of spherical inclusions (phase 2) in a matrix (phase
1)], the second-order estimates (145) for He can be simplified. The result, for
simple shear loading conditions, is given by
⎧
⎪
⎪ 1 σ20 σ20
⎪
⎨ 1 − (1 + φ 2 ) 1 − , if < 1,
σe0 2 σ10 σ10
= (185)
σ10 ⎪
⎪ σ20
⎪
⎩ 1, if ≥ 1.
σ10
An identical result is obtained for fiber-reinforced microstructures with transverse
isotropy loaded in transverse shear. We should point out that the small-contrast
expansion described in Section 4.6.1 diverges for simple shear loading, whereas,
as indicated by Eq. (185), the corresponding second-order estimate does not.
Finite-element computations carried out by Suquet (1993a), for particle-
reinforced materials with inclusion volume fraction φ2 = 0.15, indicate that,
although the two types of nonlinear estimates obtained from the linear Hashin–
Shtrikman lower bound exhibit the same general trends, the second-order estimates
are in closer agreement with the numerical results. Moreover, the variational
estimates lie above the numerical results, consistent with the fact that the vari-
ational estimates are expected to overestimate the effective yield strength of the
composite at this value of φ2 . The nonlinear estimate obtained from the linear
Hashin–Shtrikman upper bound lies below the microstructure-independent Voigt
(one-point) upper bound (see Section 4.3.1), and is such that the second-order es-
timate lies below the variational estimate, which is known to be a rigorous bound
for all statistically-isotropic microstructures.
4.7. Applications of Second-Order Exact Results 199
One may also compare the results of numerical simulations by Moulinec and
Suquet (1995) for the effective yield strength of fiber-reinforced materials with
the corresponding predictions (183) obtained from the variational method. These
authors considered cylindrical fibers (phase 2) with circular cross section and
aligned with the x3 axis, distributed randomly in a matrix (phase 1). The overall
stresses considered by these authors consisted of the superposition of uniaxial
tension and transverse shear,
= 11 (e1 ⊗ e 1 − e 2 ⊗ e2 ) + 33 e3 ⊗ e3 .
Various contrast ratios for the strengths of the two phases were investigated:
σ20 /σ10 = 0.5, 1.1, 2, 3, 5, and 10. For σ20 /σ10 = 2, 11 different realizations were
used, while for the other ratios, the computations were performed on a single
realization, representative of the average of the predictions over the entire set of
configurations for σ20 /σ10 = 2, a configuration that approaches transverse isotropy,
with its overall strain/stress response being close to the mean response of all the
realizations, both under multiaxial loading and uniaxial tension. The results are
shown in Figure 4.2. The agreement between the numerical simulation results and
the variational estimates (183) is good. In particular, the variational estimates (183)
capture rather well the flat sectors on the yield surfaces.
For the cases that involve sufficiently strong fibers, the shape of the observed ex-
tremal surfaces was found to be bimodal in character. Bimodal surfaces were used
by Hashin (1980), Dvorak and Bahei-El-Din (1987), and de Buhan and Taliercio
(1991) for describing the initial yield or the flow surface of unidirectional com-
posites. The numerical and variational results are consistent with these models and
with experimental observations (Dvorak et al., 1988). The numerical calculations
Figure 4.2. Effective yield strength 11 of composites with cylindrical fibers aligned in
the x3 -direction (perpendicular to the plane of this page) with volume φ2 . The curves are,
from left to right, for σ20 /σ10 = 0.5, 1.1, 2, 3, 5, and 10. Symbols represent the results of
numerical simulations for randomly isotropic configurations (averaged over 11 realizations),
while the curves show the predictions of the variational method in which the Hashin–
Shtrikman lower bound for the linear comparison material has been used (after Moulinec
and Suquet, 1995).
200 4. Nonlinear Rigidity and Elastic Moduli: The Continuum Approach
estimates for the effective linear properties of the comparison material with the
same microstructure as that of the nonlinear material. The material itself is a
power-law composite with the same exponent as the individual phases, and is, in
addition, incompressible. Under the assumption of statistical isotropy, the effective
potential is a function of the second and third invariant of the average strain E and,
by homogeneity, is given by Eq. (166). The variational bounds, derived above for
power-law materials provide bounds for σe0 that are independent of the parameter θ
of Eq. (166), whereas the estimates provided by the second-order theory do depend
on this parameter. It can then be shown (Ponte Castañeda and Suquet, 1998) that
the incremental and secant procedures lead to the stiffest predictions, whereas
the variational and second-order methods provide more compliant predictions.
In particular, since, as already noted in Chapter 2 (see also Chapters 4 and 7
of Volume I), the linear Hashin–Shtrikman lower bound is attained by certain
particulate microstructures, the variational estimates are actually upper bounds for
the nonlinear composites with the same type of microstructure. Therefore, both
the corresponding secant and incremental estimates violate this bound, whereas
the second-order estimates do not. In fact, the incremental estimates violate even
the Hashin–Shtrikman upper bound for statistically-isotropic microstructures, at
sufficiently large values of the exponent n. This is somewhat unexpected, as this
type of bound is known to correspond to the opposite type of microstructure, with
the stronger material occupying the matrix phase.
A similar observation was made by Gilormini (1995) in the context of the self-
consistent estimate (instead of the Hashin–Shtrikman lower bound). These results
indicate that the tendency of the incremental model to approach the Voigt (one-
point) bound (see Section 4.3.1) when m → 0 is not due to the approximate nature
of the self-consistent method, but is because of the shortcomings of the incremental
method itself. Let us emphasize again that of the four nonlinear homogenization
procedures described above, only the second-order theory yields estimates that
are exact to second order in the contrast between the properties of the phases. The
other three (variational, secant, and incremental) provide estimates that are exact
only to first order in the contrast.
Finally, Gibiansky and Torquato (1998b) derived approximations for the ef-
fective energy of d-dimensional nonlinear, isotropic, elastic dispersions. These
approximations are similar to those described in Sections 2.2.2.1 and 2.2.2.2,
derived by Gibiansky and Torquato (1998a), for the effective conductivity of the
materials with the same morphology. In addition, Gibiansky and Torquato (1998b)
derived cross-property relations that link the effective energy of nonlinear materials
with their effective conductivity.
Summary
Several continuum approaches to estimating the effective nonlinear mechanical
properties of multiphase materials were described and discussed. One method, due
to Talbot and Willis, is based on a nonlinear extension of the Hashin–Shtrikman
variational principles, while the second method, developed by Ponte Castañeda
(for nonlinear isotropic materials) and Suquet (for power-law composites) uti-
lizes new variational principles that involve a linear comparison material with the
same microstructure as that of the nonlinear composite. These methods provide
at least one type of rigorous bounds (i.e., upper or lower bounds). The Talbot–
Willis procedure yields the bounds of the Hashin–Shtrikman type, while the Ponte
Castañeda–Suquet method provides bounds and estimates of any type, given the
corresponding bounds and estimates for the linear comparison material. In both
cases, the resulting bounds and estimates are exact to first order in the contrast
between the properties of the phases. A third method, also developed by Ponte
Castañeda, yields estimates that are exact to second order in the contrast. The
resulting estimates are not, however, bounds of any type.
4.9. Critique of the Variational Procedure 203
5.0 Introduction
Beginning with this chapter, and in the next three, we study and analyze failure and
fracture of heterogeneous materials. In the present chapter, electrical and dielectric
breakdown of composite materials, which constitute a set of complex, nonlinear,
and non-local transport processes, are described. Their nonlinearity stems from
the existence of a threshold: Below and far from the threshold nothing particularly
complex happens. The laws of linear (or constitutively nonlinear) transport hold,
and the electrical properties of the materials are described by the models that were
described in the previous chapters and in Volume I. However, at the threshold,
the materials’ behavior and their transport properties abruptly change and become
very complex. Note that, unlike the percolation threshold, the threshold in electrical
or dielectric breakdown is not geometrical but dynamical although, as discussed
below, the interplay between the heterogeneities and the dynamical threshold gives
rise to a rich set of phenomena that are completely absent in the linear transport
regime in the same system.
Dielectric breakdown in gases, liquids, and solids is a complex problem and has
been studied for a long time. Many breakdown phenomena in gases are relatively
well-understood (see, for example, Meek and Craggs, 1978), while some, such as
atmospheric lightning, are more difficult to analyze, because the density, conduc-
tivity, and humidity of air are distributed inhomogeneously. Another well-known
example, in addition to lightning, is surface discharges, also known as Lichten-
berg figures. These phenomena are beyond the scope of our book and will not be
considered.
In dielectric breakdown in solids, the material is initially non-conducting when
an electric field is applied across the sample. If the field exceeds a certain threshold,
the material breaks down and becomes conducting. The microscopic mechanisms
of dielectric breakdown in solid materials are much more complex than those in
gases since, in addition to dielectric effects, mechanical and chemical effects can
also intervene and make the problem more difficult. From a practical view point,
dielectric breakdown is an important phenomenon, since it limits the application of
dielectrics as insulators. For this reason, dielectric breakdown in solids has received
much attention over the past several decades, and has been especially studied inten-
sively over the past decade.Awell-known example of such phenomena is formation
and growth of electrical trees (as in, for example, discharge treeing in polymers).
208 5. Electrical and Dielectric Breakdown: The Discrete Approach
Figure 5.1. Schematic representation of a tree growing between two electrodes on two
parallel planes (after Hill and Dissado, 1983a).
The trees themselves may not cause breakdown unless they grow so large that
they span the thickness of the material. A diagrammatic representation of this
phenomena is shown in Figure 5.1. We will come back to this phenomenon shortly.
Another important example is dielectric breakdown in metal-loaded dielectrics,
which are disordered materials consisting of a mixture of conducting and non-
conducting components. For example, solid-fuel rocket propellant is a mixture of
aluminum and perchlorate particles in a polymer binder (Kent and Rat, 1985). It has
been reported that the breakdown field of this material decreases significantly by
the presence of the aluminum particles, and is also a strong function of the volume
fraction of the constituent particles. Dielectric breakdown of such composite solids
is dominated by space charge effects due to the large electric fields near any sharp
metal tips occurring in the composite, and thus the composite is unusually sensitive
to breakdown. Recall that about a decade ago the solid fuel of a United States Air
Force rocket experienced dielectric breakdown, with the fuel becoming electrically
conductive, setting the rocket on fire.
Electrical breakdown occurs when the current through a conducting medium
causes an irreversible resistance change in the medium. In this phenomenon, the
material is initially conducting. The failure occurs when the current density flowing
in the material exceeds a threshold value at and beyond which the material becomes
insulating. Unlike dielectric breakdown, the mechanism of electrical failure is
well-understood; it is merely Joule effect which causes degradation of metallic
interconnects (or the metal lines) which, due to electromigration phenomenon,
lose their conducting properties. Note that in this phenomenon the material be-
haves precisely like a fuse, which is broken when the applied voltage exceeds a
certain limit. Electrical breakdown is a major obstacle to development of nano-
size devices. Experimental realizations of electrical and dielectric breakdown in
5.1. Continuum Models of Dielectric Breakdown 209
metal-insulator films, with a view to explain them in terms of the statistical physics
of disordered media, were reported by Yagil et al. (1992, 1993) among others. Hill
and Dissado (1983b) analyzed the older experimental data. We will come back to
these experiments later in this chapter.
Another important phenomenon that belongs to this class of problems is elec-
tromigration failure in polycrystalline metal films (see, for example, Huntington,
1975; Ho and Kwok, 1989). If a high current density passes through a thin metal
film, collisions between the conduction electrons and the metal ions result in drift-
ing of the ions and their electromigration. If there is a divergence in the flux of
the ions at some points, voids nucleate, grow and overlap with each other until
conduction ceases and the film suffers electrical breakdown (see, for example,
Rodbell et al., 1987). This phenomenon is particularly important in integrated cir-
cuits, where the continuing miniaturization of the circuits exposes the conducting
thin metal films to increasingly large current densities. Under such conditions,
electromigration failure decreases the circuit lifetime which is unacceptable from
an economical view point.
Throughout this book, both in Volume I and in the present Volume, we have
grouped the models for any phenomenon of interest to us into two classes—
the continuum models and the discrete models. In this chapter, we deviate from
this general approach because the continuum models of electrical and dielectric
breakdown of heterogeneous solid materials are well-documented (see, for ex-
ample, Whitehead, 1951; O’Dwyer, 1973; see also Niklasson, 1989a; Dissado
and Fothergill, 1992; Ohring, 1998, for more recent references); hence, the best
we could do would be summarizing these works, an unwise action. In addition,
as will be discussed in this chapter, many phenomena associated with electrical
and dielectric breakdown have a vector analogue in brittle fracture of solids, for
which many continuum models have been developed that will be described and
discussed in detail in Chapter 7. Thus, we restrict our discussion of the continuum
models to a few recent efforts that utilized extensive numerical solution of the
discretized continuum equations in order to study the breakdown phenomena in
strongly-disordered solids. On the other hand, over the past several decades several
discrete models of breakdown of heterogeneous materials have been developed.
These models are either stochastic or completely deterministic. Their general fea-
tures for modeling both the electrical and dielectric breakdown are the same, and
in fact, with appropriate modifications, a model for one of the phenomena can be
used for studying the other one. In this chapter, we describe these models in detail,
discuss their predictions, and, whenever possible, compare the predictions to the
relevant experimental data.
cE 0
V = −cE 0 cosh u cos θ + exp(β − u)( − ) sinh(2β) cos θ, u > β,
2C
(2)
cE 0
V =− (cosh β + sinh β) cosh u cos θ, u < β, (3)
C
with
C = cosh β + sinh β.
From this solution, the components of the electric field, namely, Eu = −τ −1 ∂V/∂u,
and Eθ = −τ −1 ∂V /∂θ, are computed, where τ = c(sinh2 u + sin2 θ )1/2 . One
then defines a field multiplication factor, Eu (β, 0)/E 0 = Ex (β, 0)/E 0 , which is
given by
Ex (β, 0) a( − )
= 1 + , (4)
E0 a + b
where a and b are the semi-major and semi-minor axes of the elliptical inclusion,
respectively.
The critical question to be answered is: What is the difference H in the electro-
static energy between a material with and without the inclusion? If the sources of
the applied field are fixed, then for the elliptical inclusion embedded in an infinite
medium, one has, H = − 12 px E 0 , where px is the x-component of the dipole
5.1. Continuum Models of Dielectric Breakdown 211
In analogy with brittle fracture, which is associated with a quantity referred to as the
stress-intensity factor (see Chapters 6 and 7), we define an electric field-intensity
factor KI or, more simply, field-intensity factor,
√
KI = π c E 0 ,
in terms of which one has
$ %
KI 1
Er = √ cos ,
2π r 2
$ %
KI 1
E = − √ sin , (10)
2π r 2
& $ %
2r 1
V = −KI cos .
π 2
Physically, KI is the amplitude of the r −1/2 electric field singularity at the tip of
the conducting crack. One may also define the electrostatic energy release rate
HR by
d[ 12 π c2 (E 0 )2 ]
HR = = π c(E 0 )2 , (11)
dc
where HR dc is the amount of electrostatic energy released when the crack extends
by dc, with its critical value being, HcR = 4Ec0 . Moreover,
KI c
Ec0 = √ , (12)
πc
where KI c represents the critical value of KI .
Finally, Rice (1968) developed a line integral, usually called the J -integral,
which is independent of the contour. This quantity was originally developed for
fracture of material, and its usefulness becomes evident when the contour encloses
the tip of the fracture. Thus, J yields HR , the elastic energy release rate. The
J -integral for the elasticity problem is defined by
: $ %
∂u
J = −(σ · n) · ds + He dy , (13)
∂x
where u is the displacement vector, σ is the stress tensor, n is the unit vector normal
to the contour, and He is the elastic energy density. Since the analogue of the stress
tensor is the displacement field D, then, the J -integral for the electrostatic problem
is given by
:
J = [−(D · n)Ex ds + Hdy], (14)
where H is the electrostatic energy. Garboczi (1988) showed that, similar to me-
chanical fracture, the J integral for the electrostatic energy is independent of the
path. Equation (14) was also suggested by Hoeing (1984).
5.2. Discrete Models of Electrical Breakdown 213
The above discussions should make it clear that, many results that have been
derived for brittle fracture of materials, based on the continuum models and de-
scribed in Chapter 7, can be directly translated into analogous results for dielectric
breakdown of materials.
Thus, the problem that we wish to study is the following. We are given a dis-
ordered material, represented by a lattice in which the conductance of every bond
is selected from a probability density function f (g). In this state, the material is
completely conducting (it contains no insulating region). We now select at random
a fraction 1 − p of the bonds and convert them to insulators; that is, the fraction
of the conducting bonds is p. So long as p pc , where pc is the percolation
threshold of the lattice, the material will still be conducting, albeit with a smaller
effective conductivity than when p = 1. We now apply a voltage V across the
material. If V is small enough, then there would be no change in the conductivity
state of the material. We now increase V by an amount large enough that the first
microscopic failed region (or the first failed bond in the lattice model) appears in
the material. Then, the material may behave according to one of the two scenarios.
(1) As soon as the first failed region appears, the entire material may fail rapidly
by an avalanche of local failed regions, without any need for increasing the
applied voltage V.
(2) The state of the material may be such that the macroscopic failure of the ma-
terial is more gradual, as the disorder distributes the current in an “equitable”
way, rather than concentrating it in a few weak regions. In this case, after the
first failed region appears, nothing further happens, unless we increase the
applied voltage so that new failed regions can emerge.
Corresponding to any applied voltage, there exists a current that flows through
the material. Since in practice macroscopic failure of the material is what one is
interested in, we consider the behavior of the macroscopic current and its influence
on the material. If this current exceeds a threshold If , then, the material as whole is
converted to an insulator and fails. Two important questions that must be addressed
by any model are as follows.
(1) How does If depend on p?
(2) How does the breakdown process take place? In other words, how does the
first sample-spanning path of the failed regions (or bonds in the lattice models)
appear for the first time?
Let us analyze the problem in detail for two limiting cases, namely, the
dilute limit when p 1 (very few insulating regions), and the opposite limit,
p pc (most of the sample being insulating).
and iu are, respectively, the current densities around the defect and far from it in
the unperturbed state, then, one can write
id = iu (1 + E), (15)
where E is the enhancement factor, the magnitude of which depends on the ma-
terial’s morphology. For example, for an elliptical defect with major and minor
axes 2a and 2b, E = a/b. The total current flowing through the material is then,
I = Siu = Sid /(1 + E), where S is the surface area of the electrode. The first fail-
ure happens when id = iw , where iw is the threshold current density for the failure
of the sample without the defect. Therefore,
Siw
If = , (16)
1+E
implying that the current enhancement decreases the failure current If . Typically,
the current for the complete first failure is also the current for failure of the sample,
since as soon as the regions in the vicinity of the defect fail, the current density
around the new defect is further enhanced, leading to a rapid failure of the entire
material. Clearly, the most damaging defects are those that are perpendicular to
the current lines, and are in the form of long cylinders or rods. The probability of
developing a defect depends on its shape.
In the context of the lattice models, the simplest and smallest defect is one
insulating bond which is positioned parallel to the direction of the current lines
and is far from the lattice’s boundaries (see Li and Duxbury, 1987, for the effect
of the defects that are near the boundaries of the lattice). If no defect is present in
the lattice, then, If = Liw , where L is the linear size of the lattice. For a defect of
size one (i.e., one bond), it is not difficult to show that, E = π/4, and therefore in
this case,
π
If = Liw . (17)
4
where Ld represents the volume of the system. The most probable, most damag-
ing defect is formed when PN ∼ 1, and therefore the critical number Nc for the
formation of such a defect is given by
d
Nc ∼ − ln L. (20)
ln(1 − p)
Therefore, the corresponding current density id is given by
⎧
⎪ −2 ln L
⎪
⎪ iw 1 + a 2 , 2D,
⎨ ln(1 − p)
id = (21)
⎪
⎪ −3 ln L 1/2
⎪
⎩ wi 1 + a 3 , 3D.
ln(1 − p)
Because the total current in the system is iLd−1 , the failure current is obtained by
setting id = iw , resulting in
⎧ iw L
⎪
⎪ 2D,
⎨ 1 + 2a ln L/ ln(1 − p) ,
2
If = (22)
⎪
⎪ iw L2
⎩ , 3D.
1 + 3a3 [ln L/ ln(1 − p)]1/2
The most interesting aspect of Eq. (22) is its prediction for the size-dependence
of the failure current. According to this equation, the failure current per bond,
if = If /Ld−1 , decreases with the linear size of the sample in a complex way (in
practice, L is the ratio of the linear size of the actual sample and the typical size
of the insulating defects). If ln(1 − p) is not too large, then
(ln L)−1 , 2D,
If ∼ (23)
(ln L)−1/2 , 3D.
Thus, for a fixed size of the insulating defect, the larger the sample, the smaller
the failure current.
Since the current that flows through the side link of the defect is proportional to
(⊥ )d−1 I , one obtains
(p − pc )(d−1)ν
If ∼ , (28)
(ln L)d−1
implying that, finite size of the sample generates a (weak) correction to Eq. (24).
Thermal effects also modify Eq. (24) which will be described in Section 5.2.7.
Duxbury and Li (1990) proposed that one may combine the above results for
the dilute limit and the region near pc , Eqs. (22) and (24), into a single unified
equation, given by
$ %
p − pc φ
1 − pc
If = Iw , (29)
ln(L/ξp ) ψ
1+c
ln(1 − p)
where c is a constant, and ψ is an exponent, the precise value of which is not
known, but can be bounded by
1
< ψ < 1. (30)
2(d − 1)
Thus, in general, there exist three regimes.
(1) For p = 1, one has If = Iw , as expected.
(2) For p 1, the numerator of Eq. (29) is essentially a constant of order unity,
and one recovers Eq. (22).
(3) For p pc , the denominator of Eq. (29) is of the order of 1, and one recovers
Eq. (24) with φ = (d − 1)ν.
L / If
#
$ !
voltage at which the first bond fails, and a second one, Vl , causes the last bond,
and hence the sample, to fail. The two voltages exhibit very different behaviors as
p, the fraction of the conducting bonds in the original lattice, was varied. Vi first
decreases up to p 0.7, and then increases again. On the other hand, Vl increases
monotonically with p until, in the vicinity of pc , it becomes roughly equal to Vi .
Duxbury et al. (1987) employed the same model and analyzed the dependence of
the failure current If on the sample size L. Figure 5.2 presents the results, where
L/If is plotted versus ln L. The linear dependence of L/If on ln L, for several
values of p, is in agreement with Eq. (29). In addition, when If was determined as
a function of p, it was found to follow Eq. (24) [or Eq. (29)], although when their
data are fitted to this equation, the exponent φ is about 1, rather than the theoretical
prediction (for d = 2), φ = ν = 4/3.
de Arcangelis and Herrmann (1989) studied a model of electrical breakdown in
which each conducting bond is characterized by a voltage threshold, such that if the
voltage along the bond exceeds the threshold, the bond breaks down and becomes
an insulator. This model can be thought of as the scalar analog of brittle fracture
of materials, in which a microscopic portion of a material behaves elastically
until the stress or the force that it suffers exceeds a threshold, in which case the
material breaks. The thresholds in the model of de Arcangelis and Herrmann (1989)
were distributed according to a probability distribution function. Interesting scaling
properties, in addition to what we described above, were discovered for the model.
For example, the total current I that passes through the network, as the conducting
bonds burn out, scales with the linear size L of the network as
I ∼ Lζ h(Nb /LDf ) (31)
where Nb is the number of burnt-out bonds, and h(x) is a universal scaling function.
Numerical simulations in 2D indicated that ζ 0.85 and Df 1.7. Note that
Df represents the fractal dimension of the set of all the burnt-out bonds. If one
considers only those burnt-out bonds that form a sample-spanning cluster, then
one finds that, Df 1.1, indicating that the cluster is almost like a straight line.
Moreover, de Arcangelis and Herrmann (1989) found that the distribution of the
local currents in the network just before it fails macroscopically is multifractal, so
that each of its moments is characterized by a distinct exponent (which is similar
220 5. Electrical and Dielectric Breakdown: The Discrete Approach
c k
where and are two constants, the precise values of which are not known.
The distribution (38) is similar to (35) (in the sense of being double expo-
nential) although, unlike (35), it has never been checked against the results of
computer simulations or experimental data.
(2) It is difficult to test the validity of the Gumbel distribution against the Weibull
distribution by simply fitting the data to them. However, if one defines a
quantity A by
ln[1 − FL (Vf )]
A = − ln − , (39)
Ld
222 5. Electrical and Dielectric Breakdown: The Discrete Approach
then, the corresponding quantity, for example, for the distribution (37) (when
written for the failure voltage Vf ) is given by
$ %
1
AW = a1 ln + b1 , (40)
Vf
and thus a plot of AW versus ln(1/Vf ) must be linear. On the other hand, for
the Gumbel distribution, Eq. (35) or (36) [or (38)], one has
$ %
1
AG = a2 + b2 , (41)
Vf
which predicts linear variation of AG with 1/Vf . In this way, one can clearly
determine which cumulative distribution provides a better fit of the data.
Duxbury et al. (1987) found, using this method, that the Gumbel distribu-
tion provides a more accurate fit of their numerical data. In Chapter 8 we will
utilize this method in order to test the accuracy of analogous distributions for
the failure stress of brittle materials.
mechanical failure of the first few atomic bonds generates a path of broken bonds
that eventually spans the materials), and hence we refer to this case as the “brittle”
regime. In this case, the failure of the material is governed by the weakest (or at
least one of the weakest) bonds in the initial system. For larger w, the disorder is
stronger, and therefore the breakdown of the material is more gradual, as there is
a large range over which individual bonds’ failure is driven by an increases in the
applied voltage. This situation somewhat resembles ductile fracture, and therefore
we refer to it as such (without claiming that it actually represents the scalar analogue
of ductile fracture). In a d-dimensional network of volume Ld , ductility is expected
if the number of failed bonds exceeds Ld/2 . Then, the behavior of the breaking
voltage in the ductile regime parallels that of materials with percolation disorder.
However, the average breaking voltage cannot be less than v− , and therefore this
leads to an eventual crossover to the brittle regime as the linear size L of the
network increases, except when w = 2.
Whether the network behaves as in the brittle or ductile regime depends on w
(1)
and L. Kahng et al. (1988) showed that there exists a critical value wc of w such
that, regardless of L, the material always fails in the brittle regime. The failure of
(1)
the system in this case is trivial. For w > wc , the network’s failure is brittle for
large L and ductile for small L. The two regimes are separated by another critical
(2) (2)
value of w, wc (L), which is a function of L. For L → ∞, one has wc → 2,
and failure of the system is brittle.
More quantitatively (but approximately), we consider the sequence of the weak-
est bonds. The average failure voltage for the N th weakest bond to break can be
shown to be (Kahng et al., 1988)
wN
V1 = vweak (N ) = v− + , (42)
L2
which predicts a linear dependence of V1 on N , since the distribution of the thresh-
olds is uniform and must be equal to v− for N = 0 and to v+ for N = L2 . We
now suppose that N bonds have failed and formed 2N edge bonds, where there
is an increase of the current due to enhancement effect (see above). It can then be
shown that the average failure voltage for the 2N failed bonds is given by
w
V2 = vedge (N ) = v+ + . (43)
2N + 1
Observe that V2 is a decreasing function of N, because as N increases, the proba-
bility that a weak bond is included in the 2N edge bonds increases. An approximate
criterion for brittleness of the system is then given by
EV1 (N) > V2 (N ), (44)
where E is the enhancement factor described earlier. Then, two possible situations
may arise:
(1) If we plot EV1 (N ) and V2 (N ) versus N , the two curves do not cross each other.
In this case, the network becomes unstable (behaves as in the brittle regime)
after the first bond fails, regardless of the network size L. For this to happen,
224 5. Electrical and Dielectric Breakdown: The Discrete Approach
and reaches its critical temperature faster than any other bond. The opposite limit,
b → 0, corresponds to a percolation model, because in this limit the heating rate
becomes independent of the current, and therefore the sequence in which the bonds
burn out is essentially random. Note that there are two characteristic time scales
in the system which are t1 = Tc /Ri b and t2 = 1/a. If Ic is the critical current for
the emergence of the first sample-spanning cluster of the burnt-out bonds, then,
three distinct regimes can be recognized.
(1) If the current I through the network is very close to Ic , then one has a number
of growing clusters of burnt-out bonds, all nucleating from the same center,
which is the first burnt-out bond in the network. The degree of branchiness of
the clusters depends on the quenched disorder of the network (for example,
the distribution of the resistances). The larger the disorder in the network, the
more branched the clusters are.
(2) If I Ic , then there is only one relevant time scale, t1 , in the system. Initially,
the bonds burn out more or less at random, a process that is dominated by the
quenched disorder, and then at later times the growth of the burnt-out clusters
becomes correlated as they become connected.
(3) The third regime corresponds to a crossover between (1) and (2). In this case,
the behavior of the system is extremely sensitive to the applied voltage or
current. The model produces a hierarchy of evolving failure patterns at various
length scales, as the applied current I is varied. The breakdown patterns are
also fractal with a fractal dimension Df which is a strong function of the
parameter b. Experimental realization and confirmation of this model will be
described and discussed shortly.
A stochastic model that takes into account the Joule effect was developed by Pen-
netta et al. (2000), which was intended for electrical breakdown of thin conducting
films. An external current I , which is held constant, is injected into a 2D lattice.
Each bond of the network is a resistor with a resistance r(T ) = r0 [1 + α(T − T0 )],
where r0 is a constant resistance, T is the resistor’s present temperature, T0 is a con-
stant reference temperature, and α = (1/r)dr/dT is the temperature coefficient
of resistance. A bond breaks down and becomes an insulator with a probability pb
given by
$ %
H0
pb = exp − , (49)
kB T
where ij is the current in, and N is the number of nearest neighbors of, the j th
resistor. The parameter a1 describes the heat coupling of each resistor with the
5.2. Discrete Models of Electrical Breakdown 227
substrate to which the thin film is attached, and measures the importance of Joule
heating effects. a2 is a constant which was taken to be 3/4.
Hence, starting from a resistor network in which all the bonds are conducting,
the current and temperature distributions in the network are calculated. Conducting
bonds are then converted to insulating ones with a probability given by Eq. (49).
The current and temperature distributions are then recalculated, the next bonds to
fail are identified, and so on. The simulations stop when a sample-spanning cluster
of the failed bonds is formed. Computer simulations indicated that the effective
resistance Re (t) of the sample at time t follows the following power law,
where µd 1/4. Note that the failure time tf can be estimated from two
measurements of Re (t) at two different times, namely,
ct1 − t2
tf = , (52)
c−1
where c = [Re (t1 )/Re (t2 )]1/µd represents the ratio of the two measured resistances
at two different times, raised to the power 1/µd . Therefore, once again, the concepts
of scaling and universality seem to be quite useful to modeling of an important
phenomenon, namely, electrical breakdown of thin solid films. Let us mention
that another deterministic model that takes into account the Joule effect, but uses
nonlinear, power-law, resistors (see Chapters 2 and 3) was developed by Martin
and Heaney (2000).
The second stochastic dynamical model that we describe was developed by
Hansen et al. (1990), and is a generalization of the dielectric breakdown model of
Niemeyer et al. (1984) which will be studied shortly, but also has some similarities
to the fuse model of de Arcangelis et al. (1985) described above. In their model, a
conducting bond breaks down and becomes an insulator with a probability pb ∼
η
iij , where η is a parameter of the model, and iij is the current in the bond ij .
Initially, all the bonds in the network are conducting. A macroscopic voltage drop
is applied to the network, and the current distribution in the bonds is computed.
The bond that breaks first is selected from among all the conducting bonds. The
current distribution in the network with its new configuration, including the failed
bond, is calculated, the next bond to be broken is selected, and so on.
This model provides some interesting predictions. Hansen et al. (1990) found
that there is a critical value ηc = 2 of η, such that the breakdown patterns are
qualitatively different for η < ηc and η > ηc . For η < ηc the breakdown pattern
resembles a percolation cluster, in the sense that a finite fraction of the conducting
bonds must breakdown before the system fails and becomes insulating. On the
other hand, for η > ηc the breakdown pattern is a fractal object with a fractal
dimension that depends on η. The vector analogue of Hansen et al.’s model, i.e.,
one in which the bonds represent elastic elements that break with some probability
(which might be applicable to mechanical fracture), was analyzed in detail by
Curtin and Scher (1991,1992).
228 5. Electrical and Dielectric Breakdown: The Discrete Approach
where the sum is over all the hot spots. If we assume that the resistance r and the
ratio R are the same for all the links (in the percolation material), we obtain
αr 2 R 4
B= ij (57)
4I04 j
which implies that B is related to the fourth moment of the current distribution in
the material, a subject that was discussed in Section 5.16 of Volume I. The current in
5.2. Discrete Models of Electrical Breakdown 229
pc ) = (B/B0 )
1/2 1/2(2µ+κ) , one obtains the final result:
If0
If ≥ 1/2(2µ+κ)−1/2
B 1/2(2µ+κ)−1/2 . (61)
B0
If we substitute the 2D lower bound, κ = 2ν + 1 − 2µ (see Chapter 5 of Volume I),
we obtain If ∼ (p − pc )ν , in agreement with Eqs. (24) and (28). Thus, taking the
thermal effects into account, one obtains a refinement to Eqs. (24) and (28) which
were derived earlier based on geometrical considerations alone. Since,
typically,
only a fraction of the red bonds contribute significantly to the sum i 4 , we expect
to have
1 1 1
If ∼ B −w , with − ≤w≤ . (62)
2 2(2µ + κ) 2
230 5. Electrical and Dielectric Breakdown: The Discrete Approach
We are now ready to compare the above theoretical predictions to the relevant
experimental data.
signaling the breakdown of more and more conducting fraction of the composite.
Suppose that at time tf the composite fails and becomes insulating. Lamaignere
et al. (1996) found that
tf ∼ I −2 , (63)
and that the effective conductivity ge of the composite at times close to tf follows
the power law,
ge ∼ (tf − t)µd , (64)
with µd , which is a sort of dynamical analogue of the percolation conductivity
exponent µ, being about 2/3 for their 2D material.
Yagil et al. (1992,1993) measured failure current If and the THC of thin, semi-
continuous Ag and Au percolating films. The films were evaporated in vacuum at
a rate of 0.1 nm/sec onto room temperature glass substrate. Several samples with
different surface coverage (i.e., different fraction of the conducting material) were
employed. The samples were then removed from the vacuum and measured at
room temperature. The measured I − V characteristic indicated Ohmic behavior
at low currents, and nonlinear behavior at high currents, due to Joule heating. The
failure current If was defined as the current at which the first irreversible change
in the resistance was measured. Figure 5.4 presents a sample of their results for
the failure current If versus the THC B, measured for the Ag samples.
Figure 5.4. Failure current If versus the third harmonic coefficient B, indicating the slope
w (If ∼ B −w ). The inset presents the data for the relation, B ∼ R 2+w with a slope of 3.2
(after Yagil et al., 1992).
232 5. Electrical and Dielectric Breakdown: The Discrete Approach
distribution in it is calculated. The wire that carries the most current fails first, after
which the current distribution is calculated again, the next wire to fail is identified,
and so on.
Several interesting results emerge from this model. For example, suppose that
at time t = 0 a “crack” (i.e., insulating material) of length 2c is inserted in the
metal film and its growth is monitored. Suppose also that v∞ (x, c) is the speed
of the crack tip when the crack’s length is 2x. If a constant external current flows
through the film, then for x c (Wu and Bradley, 1994)
i0 2
v∞ (x, c) x , (67)
c
where i0 is the current density far from the crack. Thus, as the crack grows, its
speed of propagation increases quadratically. The dependence
4 x on the time t of the
crack tip location for x c is obviously found from t = c dz/v∞ (z, c). Near
pc the mean failure time tf obeys the following power law
tf (p) ∼ (pc − p)ν , (68)
where ν is the exponent of percolation correlation length; clearly, tf = 0 for
p ≥ pc .
Electromigration motivates the introduction of a new percolation quantity, which
is called the minimum gap. Consider a random resistor network in which a fraction
(1 − p) of the bonds are insulating. Suppose now that a random walker starts its
walk from one side of the lattice, and jumps from one cell to an adjacent cell by
crossing the bonds, regardless of whether these bonds are conducting or insulating.
We also assume that the walk is self-avoiding, i.e., the walker never visits a cell
more than once. After some steps, the walker finally arrives at the opposite face
of the network; its path consists of all the bonds that were visited. Suppose then
that the path consists of Nc conducting and Ni insulating bonds. The connection
between this concept and electromigration becomes clear if we assume that, any
bond that is crossed by the walker breaks down and becomes an insulator. Thus,
in a 2D system, for example, when the walker has crossed the sample, the system
breaks down and becomes an insulator. The shortest path is one that corresponds
to the smallest number of resistors that burn out during the walk.
We now introduce the concept of minimum gap gm which, in an insulating
material, is the minimum number of conducting bonds (per length of the system)
that must be added to the system (or to the trail of the random walk) in order for
the material to become conducting. Clearly, gm depends on p, the fraction of the
conducting bonds already in the material. Chayes et al. (1986) and Stinchcombe
et al. (1986) studied the properties of the minimum gap gm (p). Figure 5.5 present
the dependence of gm (p) on p in the square network. For p 1, the minimum
gap decreases from 1, with the slope dgm /dp 3 in the square network. Near pc ,
the minimum gap vanishes according to the power law,
gm ∼ (p − pc )ν . (69)
Thus, Eqs. (68) and (69) suggest that the failure time is proportional to the mini-
234 5. Electrical and Dielectric Breakdown: The Discrete Approach
Figure 5.5. Dependence of the minimum gap gm , normalized by the linear size of the
square lattice, on the fraction p of the conducting bonds (after Manna and Chakrabarti,
1987).
mum gap gm (p) of the network. On an intuitive ground, the relation between the
minimum gap and the time to failure in the electromigration problem is expected.
A problem related to electromigration phenomenon is one in which the line
width of the metallic interconnects is comparable to, or smaller than, the grain size
of the film. In this case, referred to as the bamboo regime, the grain boundaries
no longer provide connected diffusion paths along the conductor line. Instead,
electrical breakdown occurs due to intergranular voids which nucleate at the edges
of the line, migrate in the current direction, and finally collapse into a slit which
disconnects the conductor.
This problem was studied in detail by Schimschak and Krug (1998), and later
by Mahadevan et al. (1999), whose analysis we briefly describe. The shape of the
void changes due to the current I along its inner surface. Two factors contribute
to the current, the electromigration and capillary smoothing. Thus, one writes
∂Y(L)
I =γ σ + qE(L) , (70)
∂L
where γ and σ are, respectively, the atom mobility and the surface tension, L is
the arc length along the surface, Y is the surface curvature, q is the charge, and E
is the tangential local electric field. Because of conservation of the void area (in
5.4. Dielectric Breakdown 235
2D), the inner surface must move with a normal velocity vn which is given by
∂I
vn + = 0. (71)
∂L
Due to the growth of the void, this is a moving boundary-value problem, the
numerical solution of which is typically difficult to obtain.
One must first determine the electric field E by solving the Laplace’s equation
in the domain outside the void, subject to the boundary conditions that the normal
electric field vanishes at the void surface, and a constant electric force E0 is applied
to the system far from the void. It is√not difficult to see that the only relevant
length scale in the problem is s = σ/(qE0 ), and therefore the natural time
scale is given by, ts = 4s /(σ γ ), with which the governing equations can be made
dimensionless. After determining the distribution of the electric field, Eq. (71) is
iterated. A breakup procedure is triggered if two points that belong to different
surface segments are closer than half the distance between neighboring points
along the surface. In a similar way, merging of two voids can be treated.
Numerical simulations of this model indicated that, typically, the void disin-
tegrates at long times by one of the two routes. If the void is initially elongated
along the current direction, then, a protrusion develops at the leading end of the
void, which subsequently forms a daughter void. Because the daughter void is
smaller than the initial void, it moves more rapidly ahead of the mother void. If,
on the other hand, the void is initially deformed perpendicular to the current, an
invagination develops which eventually splits the void horizontally.
Consider now the dual of the 2D material which is obtained by replacing the
conducting phase by the insulating material and vice versa. We also assume that
the conductivity g of the formerly-insulating parts is given by, g = 1/. The dual
material is conducting since the original material was assumed to be insulating or
dielectric, and therefore the current I must satisfy the continuity equation, ∇ · I =
0, because of which one can write, I = ∇ × ψ, where the potential vector ψ is
selected such that only its z-component ψz (x, y) = 0. As I = g(r)E, we must have
∂ 1 ∂ψz ∂ 1 ∂ψz
+ = 0, (74)
∂x g(r) ∂x ∂y g(r) ∂y
or,
∂ ∂ψz ∂ ∂ψz
(r) + (r) = 0. (75)
∂x ∂x ∂y ∂y
In view of Eq. (73), we see that the conductivity problem in the dual material is
identical with the dielectric problem in the original composite, if ∂ψz /∂x = ∂φ/∂x
and ∂ψz /∂y = ∂φ/∂y. If so, one has, Ix = ∂ψz /∂y = Ey and Iy = −∂ψz /∂x =
−Ex . Therefore, the magnitude of the current density I in the dual material is
equal to that of E in the original composite, but its direction is rotated by 90◦ from
the dielectric problem. Physically, while in the electrical breakdown problem the
current is zero inside an insulating inclusion, in the dielectric breakdown problem
the electric field is zero inside a conducting region. Moreover, the regions that
experience an enhancement of the current (in the electrical breakdown problem)
are perpendicular to those that feel the enhancement of the electric field (in the
dielectric breakdown problem). The conclusion is that, in 2D, most of the results
that were described above for the electrical breakdown problem can be immediately
translated to corresponding predictions for the dielectric breakdown problem. We
will discuss this important point shortly, but let us first describe discrete models
of dielectric breakdown.
However, Niemeyer et al.’s model does not have an explicit rule for breakdown.
A bond with even a small probability pb can break down, which is not realistic.
Moreover, the physical reason for Eq. (76) is not clear. Pietronero and Wiesmann
(1988) did attempt to give a theoretical justification for Eq. (76) based on the
time required for the establishment of a filamentary projection of the discharge
as a sort of a “conducting fluid” in a given region of the local field. While their
argument may justify use of Eq. (76), in the limit η = 1, for dielectric patterns
in gases, its generality is not clear, and in addition, whereas the structure of the
simulated discharge patterns is highly sensitive to η (Barclay et al., 1990; Sánchez
et al., 1992), the physical origin or significance of η is not clear. Moreover, the
breakdown patterns in solid materials are propagating damage structures, not the
advancing front of an injected charge “fluid,” as in Niemeyer et al.’s model. As
such, their model is not, in general, suitable for dielectric breakdown in solids.
Wiesmann and Zeller (1986) (see also Noskov et al., 1995) modified the η-
model by incorporating two new features in it. One was that a critical field Vc
for the growth of the dielectric pattern was introduced, such that the breakdown
probability pb is non-zero if Vij ≥ Vc , and pb = 0 otherwise, an assumption that
makes the model somewhat similar to the deterministic models discussed in the
next section. The second feature was the introduction of an internal field Vs in
the structure, such that the potential in it is no longer V0 but V0 + sVs , where s
is the length of the path (measured as the number of sites that it contains) along
the structure which connects the point to the central electrode. The structure of
the resulting dielectric pattern now depends on Vc and Vs . Figure 5.6 shows two
of the fractal patterns generated by this model which are somewhat similar to
treeing in polymers. However, the accumulation of damage, which is known to be
required for electrical tree formation in AC fields, is not allowed in the Wiesmann–
Zeller model, and therefore their model is probably more appropriate for nano-
Figure 5.6. Dielectric trees with the ground plate and the needle voltage V = 0 and the top
plate at V = V0 . The threshold field for growth is zero for the left pattern, and about the
original field at the tip for the right pattern (after Wiesmann and Zaller, 1986).
5.4. Dielectric Breakdown 239
second impulses. Even then the damage pattern situation is not fractal (Knaur and
Budenstein, 1980), whereas the Wiesmann–Zeller model predicts it to be fractal.
Dissado and Sweeney (1993) argued that fractal tree-like patterns should form only
when the fields at the growth tips can fluctuate around their values obtained from
the solution of the Laplace’s equation. They showed that if one treats the local-field
enhancement factor as a white noise generated by the breakdown mechanism itself,
the amount of branching in the dielectric pattern would depend only on the range
of the fluctuations allowed. Thus, the Wiesmann–Zeller model, though interesting,
is not also completely suitable for modeling dielectric breakdown in solids.
0.10
0.05
0.00
0.8 0.9 1.0
p/pc
η
is the largest value of BVij among all the interface bonds. The second model is
clearly very similar to the model of Niemeyer et al. (1984). Breakdown patterns
were found to be fractal again, with a fractal dimension that depended sensitively
on η. In the model of Manna and Chakrabarti (1987), each bond or site of the lattice
is either conducting with probability p or insulating (dielectric) with probability
1 − p. After determining the voltage distribution throughout the lattice, all the
insulating bonds or sites break down if the voltage that they suffer is larger than
a threshold voltage. Chakrabarti et al. (1987) and Barbosa and de Queiroz (1989)
studied this model with small-cell position-space renormalization group approach.
Bowman and Stroud (1989) studied the same model, except that in their work the
insulating bond with the largest voltage difference between its end sites breaks
down first. In a somewhat different model, Benguigui (1988) considered the case
in which after a bond breaks down it becomes a superconductor. This was achieved
by inserting light emitting diodes as the insulators in a host of conductors.
The most critical questions in dielectric breakdown phenomenon, that any
reasonable model should be able to address, are as follows.
(1) How does the initial breakdown field Eb (or the corresponding voltage Vi )
depend on the volume fraction p of the conducting material (bonds) in the
initial dielectric material? A typical example is shown in Figure 5.7.
(2) How does the final voltage Vf vary with p? For small p one expects the
final breakdown voltage Vb = Eb L to be different from the initial breakdown
voltage, but as p increases the difference between the two decreases until very
near pc where they are essentially identical. This has an important consequence
in that, when these two voltages are equal, the breakdown proceeds by an
avalanche (see the discussion above) in that, many bonds break down without
any need for further increase in the applied macroscopic voltage drop.
(3) How do the two voltages Vi and Vb depend on the linear size L of the sample?
To understand the importance of the sample size, recall that breakdown starts
5.4. Dielectric Breakdown 241
near the critical defect of the system, which is (roughly speaking) the largest
pair of strongly interacting conducting clusters which are oriented parallel
to the macroscopic electric field. The breakdown field is of the order of the
inverse of the linear size of the defect, and since the largest defect in a large
system is larger than the largest defect in a small sample, the breakdown field
is smaller in the larger sample.
(4) How does the path length, i.e., the number of bonds in the breakdown path,
vary with p? An example is shown in Figure 5.8.
(5) Do power laws govern the important properties of dielectric breakdown (such
as the breakdown field Eb ) near pc , and if so, are such laws universal?
We now discuss the scaling laws that govern the dependence on p of various
properties of interest near the percolation threshold, and also on the sample size L.
only are such scaling properties important, but are in fact measured routinely in
practical situations, and therefore a scaling theory of breakdown phenomena in
terms of the sample size L is a very useful tool for interpreting the experimental
data.
The scaling properties of dielectric breakdown phenomena have been studied
extensively. Let us first recall that, as discussed in Chapter 6 of Volume I, the static
dielectric constant 0 follows the following power law (Efros and Shklovskii,
1976) near the percolation threshold pc ,
To estimate y, Beale and Duxbury (1988) used an argument based on the idea of
the critical defect mentioned above. Suppose that the total length of the critical
defect (the conducting path), made up of a pair of the largest interacting clusters of
conducting material, separated by a small distance, is . The electric field between
these two clusters is enhanced by a factor of the order of times the applied
macroscopic field. Far from pc the probability of finding a percolation cluster of
linear size is given by Eq. (26). The largest cluster in a d-dimensional percolation
system of volume Ld is of the order of m ∼ ξp ln Ld [see Eq. (27)]. Since Eb ∼
1/m , we obtain (Beale and Duxbury, 1988)
(pc − p)ν
Eb ∼ , (79)
ln L
and therefore y = ν, which is certainly greater than s/2. Equation (79) can also
be derived based on the argument (Stinchcombe et al., 1986) that Eb should be
proportional to the minimum gap gm which is proportional to ξp−1 . The ln L term of
Eq. (79) can also be derived from the fact that (Li and Duxbury, 1987) the maximum
current Im in a percolation network of linear size L that leads to its failure is given
by Im ∼ (ln L)ψ , where ψ is the same exponent that appears in (29) and (30)
(for the problem of the largest currents in a random resistor network see also
Machta and Guyer, 1987). Numerical simulations (Manna and Chakrabarti, 1987;
Benguigui, 1988; Beale and Duxbury, 1988; Bowman and Stroud, 1989) seem to
confirm Eq. (79). Lobb et al. (1987) and Chakrabarti et al. (1988) extended this
analysis to the Swiss-cheese model of continuum percolation, in which spherical
or circular grains of dielectric are distributed randomly in a conducting matrix,
5.4. Dielectric Breakdown 243
system with percolation-type disorder, such as the Swiss-cheese model, Eq. (86)
is no longer valid. Instead, one has a simple exponential, Weibull-like distribution.
Summary
Discrete models of electrical and dielectric breakdown of composite solids have
provided very useful insights into the properties of these important phenomena, by
demonstrating the significant role that defects of heterogeneities play in them. In
particular, they have provided the important prediction that the statistics of these
breakdown phenomena depend critically on the volume fraction of the defects or
the broadness of the distribution of the heterogeneities. If the volume fraction of
the defects is low, then, the probability distribution of the failure fields (voltage or
current) is of Gumble type, rather than the classical Weibull distribution. Moreover,
the discrete models have enabled us to obtain important predictions for the effect
of sample size on breakdown properties of heterogeneous materials.
6
Fracture: Basic Concepts and
Experimental Techniques
6.0 Introduction
In Chapter 5 we studied electrical and dielectric breakdown of materials—
phenomena that are well-known examples of nonlinear scalar transport processes
with their nonlinearity manifested by the existence of a threshold in the linear
(or possibly nonlinear) constitutive law that describes the relation between the
flux and the potential gradient. Beginning with this chapter, we study a nonlinear
vector transport process which is of immense significance to materials, and leads
to their mechanical failure. This type of failure, which is a result of nucleation
and propagation of fractures in materials, varies anywhere from brittle fracture,
that represents a nonlinear vector transport process characterized by a threshold
in the otherwise linear elasticity equations that govern the elastic behavior of the
material, to ductile yielding and flow. Such failure phenomena are some of the
most complex sets of phenomena in science and technology. The range of natural
and industrial systems in which mechanical fracture occurs is very broad. Under
a large stress or strain, a crack opens up in soils which grows with time, lead-
ing to complex phenomena such as soil liquefaction and eventually earthquake.
Natural or man-made fractures in oil and geothermal reservoirs and aquifers are
crucial to the flow of oil, heat and vapor, or groundwater, especially in those reser-
voirs that have a very small porosity, such as many oil fields in the Middle East.
Other rock-like materials, such as concrete and asphaltenes, often develop large
fractures, causing considerable damage to highways and buildings. Propagation of
cracks in airplane wings and fuselages can cause an airliner to crash. An important,
and undesirable, property of many high-temperature superconducting materials is
their brittleness and mechanical instability. Polymers, glasses and ceramics often
develop microcracks under a large enough stress or strain which can lead to their
mechanical failure and eventual fragmentation. Composite materials can develop
cracks due to thermal mismatch between their various constituents. Pressurized
nuclear reactors can develop cracks in their structure which can create tremendous
safety problems. Thus, a comprehensive understanding of fracture nucleation and
propagation has tremendous practical implications.
248 6. Fracture: Basic Concepts and Experimental Techniques
1According to legends, Humpty Dumpty was a powerful cannon that was mounted on top of St.
Mary’s at the Wall Church in Colchester, defending the city against siege in the summer of 1648, during
the English Civil War (1642–1649). The church tower was hit by the enemy, with its top blown off,
hence sending Humpty to the ground.
6.1. Historical Background 249
governing the movement of projectiles. As we now know, these two subjects have
fared differently over the centuries. The first subject, now known as the strength of
material, is an integral part of the basic education that most engineering students
receive, while the second one has become a core subject that physicists learn at
the beginning of their education. Although now, as in Galileo’s time, shipbuilders
need good answers to questions about the strength of materials, the subject has
never yielded easily to basic analysis. Galileo identified the main difficulty when
he wrote: One cannot reason from the small to the large, because many mechanical
devices succeed on a small scale that cannot exist in great size. Over 350 years
after Galileo wrote these lines science reached the atomic scale and began to
answer the questions that he had posed on the origins of strength and the relation
between large and small. These wise words of Galileo also pointed out an important
aspect of fracture of materials, namely, the fact that this is an inherently multiscale
phenomenon, ranging from atomic to macroscopic length scales. While the vast
majority of the theoretical and computer simulation studies of fracture have been
concerned with only one of these length scales, the past few years have witnessed
development of multiscale modeling approaches to fracture propagation in solid
materials. We will describe such approaches in Chapter 10.
However, huge accidents in the 1800s and the first half of the twentieth century,
that were caused by catastrophic fracture of materials, provided the motivation for
intensive study of fracture phenomena. For example, the boiler of the Soltana, a
steamboat that carried the Union soldiers during the American Civil War, exploded,
resulting in the death of over 1,000 soldiers. In 1919, a molasses tank 50 feet high
and 90 feet wide burst in Boston, killing 12 people and several horses. The court
auditor concluded that, the only rock to which he could safely cling was the obvious
fact that at least one-half of the scientists must be wrong.
One of the most important cases of material fracture in the twentieth century,
that helped to establish the significance of fracture mechanics, occurred during
World War II. Wartime demands for ocean freighters led to the production of the
Liberty ship, the first to have an all-welded hull. Of the nearly 4,700 ships of the
Liberty class launched during the war, over 200 suffered catastrophic failure, some
splitting in two while lying at anchor in port, and over 1,200 suffered some sort
of severe damage due to fractures. The discipline of fracture mechanics emerged
from these catastrophes. The all-welded ships were redesigned, eliminating, for
example, sharp corners on hatches, and systematic procedures were developed for
testing the fracture resistance of materials. In the early 1950s, failure by fracture
cursed the British airline industry’s efforts to establish passenger service using
jet aircraft. Ill-placed rivet holes destroyed two of Britain’s Comet aircraft, and
played an important role in transferring the center of gravity for building civilian
jet aircrafts from Britain to the United States. Aircrafts are now subjected to a
systematic program of inspection that acknowledges that every structure has flaws,
but that flaws greater than a certain size are intolerable. Testing procedures have
continued to evolve in response to accidents, most recently after an incident (in
the 1980s) in which part of the top of the fuselage of an Aloha airliner separated
during flight, killing two people.
250 6. Fracture: Basic Concepts and Experimental Techniques
the crack tip, causing the tip to become blunt. Brittleness and ductility are not,
in fact, inherent in the atoms that make up a solid. For most solid materials there
is a definite temperature at which they make a transition from brittle to ductile
behavior which for example, is about 500◦ C for silicon. In Chapter 7 we will
briefly describe theories that attempt to predict this transition temperature.
6.5.5 Crazing
Crack nucleation by crazing occurs in amorphous polymeric materials. When such
materials are deformed by an applied stress, the polymeric chains rotate and, if the
strain is large enough, become aligned in the direction of the maximum extensional
strain. Crazing then involves formation of planar arrays of fine voids that are normal
to the tensile stress. The distance between the voids is filled by ligaments of aligned
polymer chains. If the deformation is strong enough, the ligaments eventually break
and help the voids to merge.
For these reasons, Mode I corresponds most closely to the conditions used in
most experimental and theoretical work on brittle fracture of solids, since there
is always a tendency for a brittle crack to seek an orientation that minimizes the
shear loading. This is consistent with crack extension by progressive stretching
and rupture of cohesive bonds across the crack plane. In 2D isotropic materials,
Mode II fracture cannot easily be observed, because slowly propagating fractures
spontaneously orient themselves so as to make the Mode II component of the
loading vanish near the crack tip (Cotterell and Rice, 1980). Mode II fracture is,
however, observed in strongly anisotropic materials. For example, friction and
earthquakes along a pre-defined fault are examples of Mode II fracture where the
binding across the fracture interface is considerably weaker than the strength of
the bulk of the material. Pure Mode III fracture, although experimentally difficult
to achieve, is sometimes used as a model system for theoretical studies, since
in this case the equations of elasticity simplify considerably. Analytical solutions
obtained in this mode (some of which will be described in Chapter 7) have provided
considerable insight into the fracture process.
&
σ?
$ 5JHAII
"
σOO
> σNN
ρ )
respectively. For over 80 years, Eq. (7) has remained a pillar of the classical
continuum theory of brittle fracture.
To develop his theory further, Griffith took advantage of the Inglis’ solution
for an elliptical cavity described above. It can be shown that for a system under
constant applied stress (during crack formation), HM = −HE , where HE is the
strain potential energy stored in the elastic material, mentioned above, and the
negative sign is due to the fact that crack formation reduces the mechanical energy.
Using the solution of Inglis, it is not difficult to compute the strain energy density,
from which one obtains (by integrating the energy density over dimensions that
are large compared with the length of the crack), HE = −HM = π c2 (σ 0 )2 /Y ,
where Y is equal to the Young’s modulus Y in plane stress (thin plates), and
Y = Y /(1 − νp2 ) in plane strain (thick plates), with νp being the Poisson’s ratio.
Since, for a unit width of the crack front, one has HS = 4c, where is the free
surface energy per unit area, one obtains
π c2 (σ 0 )2
H = 4c − . (8)
Y
If we now apply Griffith’s criterion, Eq. (7), and identify σ 0 = σc0 as the critical
stress, we obtain
&
2Y
σc =
0
. (9)
πc
Equation (9) is the famous Griffith relation, and is the mechanical analogue of
Eq. (5.8), the critical value of the far-field electric field for dielectric break-
down. Griffith also succeeded in qualitative verification of Eq. (9) by carrying
out experiments on an inorganic glass.
Because d 2 H/dc2 < 0, the energy of the system at equilibrium is maximum,
and therefore its configuration is unstable. That is, for σ 0 < σc0 the crack remains
stationary at its initial size c, whereas for σ 0 > σc0 it propagates spontaneously
without limit. Note, however, that an unstable crack may ultimately be arrested at
some point, which is often the case with cracks around contacts and inclusions. In
this case, further increase in the applied loading may lead to a second, catastrophic
instability configuration.
Kβ which lead to a stress field that is a linear combination of the pure modes:
3
Kβ β
σij = √ fij (v, θ ). (15)
β=1
2π r
The critical condition for crack propagation can now be expressed in terms of
the critical value Kc of the stress intensity factor, which is usually referred to as the
fracture toughness. Thus, in terms of the critical energy Hc , the fracture toughness
is given by
!
Kc = Hc Y . (16)
We should emphasize, as already mentioned above, that the Griffith–Irwin pre-
diction for the critical stress σc0 (or the fracture toughness Kc ) is valid for the onset
of growth under static conditions, and for homogeneous materials. As soon as the
crack begins to grow, the stress field around it changes dynamically. In particular,
if the crack propagates at high speeds, the inertial effects substantially change the
stress field. The Griffith–Irwin approach has nothing to offer for these changes.
In other words, the Griffith–Irwin criterion can tell us when a brittle crack may
extend, but has nothing to say about how it will extend. In addition, the r −1/2 sin-
gularity at the tip of the crack cannot be reconciled with any real fracture process,
as there is no solid that can resist an infinite stress anywhere in its structure. The
root of this singularity is in the assumptions that the Hooke’s law (linear elasticity)
is operative everywhere in the material, and that a continuum approximation can
describe the state of the system. These assumptions break down for the region in
the vicinity of the crack tip, and necessitate a reclassification of the region around
the tip; this is discussed in the next section.
separate regions:
(1) The cohesive zone, which is the region immediately surrounding the crack tip,
in which all the nonlinear dissipative processes that allow a crack to move
(forward) are assumed to occur. In continuum fracture mechanics, detailed
description of this zone is avoided, and is simply characterized by the energy
, per unit area of crack extension, that it will consume. The size of the
cohesive zone is material-dependent, ranging from nanometers in glass to
microns in brittle polymers. Its typical size is the radius at which an assumed
linear elastic stress field surrounding the fracture tip would equal the yield
stress of the material.
(2) The universal elastic region, which is the region outside of the cohesive zone
for which the response of the material can be described by linear continuum
elasticity. Outside of the cohesive zone, but in the vicinity of the fracture tip,
the stress and strain fields take on universal singular forms which depend only
on the symmetry of the externally applied loads. In 2D the singular fields
surrounding the cohesive zone are completely described by the three stress
intensity factors which incorporate all the information regarding the loading
of the material. As discussed above, the stress intensity factors are related to the
energy flux into the cohesive zone. The larger the overall size of the material
containing the crack, the larger this region becomes. Roughly speaking, for
given values of √ the stress intensity factors, the size of the universal elastic
region scales as L, where L is the macroscopic length scale on which forces
are applied to the material. Thus, as L increases, the assumptions of continuum
fracture mechanics become progressively more accurate.
(3) Outer elastic region, which is the region far from the crack tip in which stresses
and strains are described by linear elasticity. Details of the solution of the
equations, describing fracture propagation, in this region depend only on the
locations and strengths of the loads, and the shape of the material. For some
special cases, analytical solutions are available, but in general one must resort
to numerical simulation. That deriving these solutions is possible is because, so
far as linear elasticity is concerned, viewed on macroscopic scales, the cohesive
zone shrinks to a point at the fracture tip, and the fracture itself becomes a
branch cut. Thus, replacing the complex domain in which linear elasticity holds
with an approximate one that needs no detailed knowledge of the cohesive zone
is another approximation that becomes increasingly accurate as the dimensions
of the sample, and hence the size of the universal elastic region, increase. The
assumption that the cohesive zone in a material is encompassed within the
universal elastic region is sometimes called the assumption of small-scale
yielding.
The dissipative processes within the cohesive zone determine the fracture energy
. If no dissipative processes other than the direct breaking of the atomic bonds
take place, then will be a constant that depends on the bond energy. In general
though, is a complex function of both the crack velocity and history, and differs
by orders of magnitude from the surface energy—the amount of energy required
6.10. Dynamic Fracture 263
the crystalline structure of the glass is relatively open. Once inside the glass, water
forms a base with existing sodium ions which corrodes the region in the vicinity of
the crack tip, hence lowering its toughness and increasing the likelihood of brittle
fracture. Aluminum and titanium, two heavily-used metals in aircraft, suffer most
from stress corrosion cracking.
The mechanism that leads to stress corrosion cracking is either anodic or ca-
thodic. That is, the phenomenon can be suppressed in an electrolytic environment
by placing either the anode or the cathode on the material and the corroding agent
as electrolytic medium. For example, hydrogen embrittlement of metals is the
most common cathodic process. The anodic process also occurs in metals that are
coated by a layer of oxide to be protected from the environment. If the coating
is opened at the crack tip, the metal will be exposed to the anodic agent at the
tip. Under this condition, the velocity of fracture propagation would be controlled
by the rate of the chemical reactions. Since these reactions are typically slow, the
fracture propagates slowly, which is why, for example, it takes aircraft a long time
to develop stress corrosion cracks in their fuselage.
factor (or stored energy density) along the prospective path of a crack can in-
crease, resulting in a continuously accelerating crack, or decrease, leading to a
decelerating and possibly arrested crack. A few examples of the common loading
conditions that are used are shown in Figure 6.5 where single-edge notched (SEN),
double-cantilever beam (DCB), and infinite strip (IS) loading conditions are pre-
sented. The SEN condition is sometimes used to approximate fracture propagation
in a semi-infinite system. When the external loading is a constant stress applied
at the vertical boundaries of the sample,√then for a large enough sample the stress
intensity factor KI is proportional to σ l, and therefore the energy release rate H
is given by, H ∝ σ 2 l (see Chapter 7 for additional theoretical details), where σ is
the applied stress and l is the length of the crack. This configuration is used, for
example, to study the behavior of an accelerating crack.
In the IS configuration, the sample is loaded by displacing its vertical boundaries
by a constant amount. Under this condition, the energy release rate is constant for
a crack that is sufficiently far from the horizontal boundaries of the sample, and
thus this loading configuration is amenable to the study of a crack moving in
steady-state.
In the DCB configuration, a constant separation of the crack faces is imposed
at l = 0, H ∝ l −4 is a decreasing function of l, and hence can be used to cause
crack arrest. How is the DCB configuration used to study dynamic fracture? An
initially imposed seed crack of length l = l0 would propagate as soon as H ex-
ceeds the limit imposed by the Griffith condition, i.e., when dH/dl = 0. Under
ideal conditions, the crack propagates for an infinitesimal distance and then stops,
because in DCB configuration H is a decreasing function of l. Although the Grif-
fith criterion assumes that the initial crack is as sharp as possible, what is prepared
in the laboratory by cutting rarely yields a tip that meets this condition. We may
view the initial seed crack as having a finite radius at its tip, thereby blunting the
stress singularity and allowing a substantially higher energy density to be imposed
in the system prior to fracture than what is allowed by a sharp crack. Thus, there
is excess elastic energy that drives the crack beyond the constraints imposed by
an initially sharp crack which, in the case of the DCB configuration with constant
266 6. Fracture: Basic Concepts and Experimental Techniques
separation imposed, can cause a crack to propagate well into the sample before
crack arrest occurs. Nonlinear material deformation around the tip, plastic flow
induced by the large stress build-up, and from crack-tip shielding that results from
the formation of either micro-cracks or small bridges across the crack faces in the
near vicinity of the tip, can also cause blunting of the singularity around a crack
tip. The DCB configuration can also generate an accelerating crack by imposing
a constant stress (instead of constant separation) at the crack faces. Under this
condition, the (quasi-static) energy release rate increases with the crack length l as
12σ 2 l 2
H= , (18)
Y w2 d 3
where σ is the stress applied at opposite points on the crack faces at the edge of the
sample, and w and d are, respectively, the thickness and half-width of the sample.
6.11.2.2 Photoelasticity
This method, coupled with high-speed photography, is also used for measuring the
stress distribution, and hence the stress intensity factor, induced by a moving crack
(Kobayashi, 1987). It is based on the birefringence induced in most materials un-
der an imposed stress, which causes the rotation of the plane of polarization light
moving through the material. The induced polarization depends on the properties
of the stress tensor which are rotationally invariant, and therefore can depend only
on the two principal stresses σ1 and σ2 . Moreover, there should be no rotation of
polarization when the material is stretched uniformly in all directions, in which
case the two principal stresses are equal, and therefore the angular rotation of the
plane of polarization must be of the form, c(σ1 − σ2 ), where c is a constant that is
determined experimentally. If stresses of a 2D problem are calculated analytically,
the results can be substituted into this expression and compared with experimental
fringe patterns obtained by viewing a reflected or transmitted beam of incident
polarized light through a polarizer. The observed intensity depends on the phase
268 6. Fracture: Basic Concepts and Experimental Techniques
difference picked up while traversing the material, and hence provides a quantita-
tive measure of the local value of the stress field. The application of this method to
transparent materials is straightforward. These methods have also been extended
to opaque materials by the use of birefringent coatings which, when sufficiently
thin, mirror the stress field at the surface of the underlying material. Dally (1987)
reviewed the applications of these methods. Similar to the method of caustics,
quantitative interpretation of these measurements is limited to the region outside
of the plastic zone.
velocity over a short, say about 3 µs, interval. Moreover, precision of this method
is obviously limited by the accuracy at which the location of the crack tip can be
determined from a photograph.
These problems can, to some extent, be overcome by using a streak camera
(Bergkvist, 1974). In this method, a film is pulled past the camera’s aperture at
high speed. The material is illuminated from behind so that, at a given instant, only
the light passing through the crack is photographed. Since one can force a crack to
propagate along an essentially straight line, the exposed film provides a continuous
record of its length as a function of time. The basic resolution of the measurements
depends on the film’s velocity and that of the high-speed film used, and on the
post-processing performed on the film in order to extract the velocity measurement
and the stability of the film’s travel velocity. The same type of experiments have
also been carried out (Döll, 1975) by high-speed measurements of the total beam
intensity that penetrates the material. If the crack does not change its shape, the
beam intensity depends linearly on the crack’s length.
as it traverses the sample, the trace of which is imprinted onto the resulting fracture
surface. Since the temporal frequency of the modulation is that of the ultrasonic
driving, measuring the distance between neighboring surface modulations pro-
vides a nearly continuous data set for the instantaneous velocity of the crack tip.
The method’s precision is limited only by the ultrasonic frequency used, which
is typically in the MHz range, and also by the precision of the surface measure-
ment. The disadvantage of this method, relative to the other techniques, is that it
is a perturbative method, since the crack deflection is accomplished by altering
the stress field at its tip, and hence externally-induced oscillations can potentially
mask intrinsic, time-dependent effects.
Langford et al., 1989, and references therein), and that periodic stress waves are
emitted from the tip of the rapidly moving cracks in a wide variety of materials
(see, for example, Rosakis and Zehnder, 1985; Dally et al., 1985, and references
therein). Fineberg et al. (1991,1992) carried out beautiful and precise experiments
to study fracture propagation in brittle plastic PMMA and showed, that there is
a critical velocity vc beyond which the velocity of crack tip begins to oscillate,
the dynamics of the crack changes abruptly, and a periodic fracture pattern is
formed. For v > vc the amplitude of the oscillations depends linearly on the mean
velocity of the propagating crack. Thus, the dynamics of cracks is governed by a
dynamical instability, and explains why the crack tip velocity does not attain the
limiting Rayleigh velocity predicted by the linear elastic theory. Although Yoffe
(1951) had already predicted the existence of a sort of dynamical instability in
fracture, showing that a fracture that moves along a straight line will branch off if
its speed becomes larger than a critical value, her predicted critical velocity was
too large, and therefore the type of instability that was considered by her could not
provide a complete explanation for Fineberg et al.’s experiments. The theoretical
studies of such fracture patterns will be discussed in Chapter 7.
In another set of beautiful experiments, Gross et al. (1993) used two materials,
the PMMAand soda-lime glass, to show that all features of dynamics of crack prop-
agation in the two materials, such as acoustic emission, crack velocity, and surface
structure, exhibit quantitative similarity with each other. Thus, there exists univer-
sal characteristics of fracture energy in most materials that are the result of energy
dissipation in a dynamical instability. Perhaps the most spectacular experiments
were carried out by Sharon et al. (1995) and Sharon and Fineberg (1996) using
the brittle plastic PMMA. They identified the origin of the dynamical instability
6.13. Mirror, Mist, and Hackle Pattern on a Fracture Surface 273
during fracture propagation as being the nucleation and growth of the daughter
cracks which limit the speed of the propagating crack tip. The daughter fracture
carries away a fraction of the energy concentrated at the tip of the moving crack,
thus lowering the velocity of the tip. After some time, the daughter crack stops
growing, and thus the crack tip velocity increases, until a new daughter fracture
starts to grow, and so on. They also observed that the branching angle for a longer
daughter fracture was smaller than that of the shorter daughter fractures. Theoret-
ical modeling and computer simulations of dynamic fracture that can reproduce
these features will be described in detail in Chapters 7 and 8.
1 mm
Figure 6.7. Light microscope photograph of mirror, mist, and hackle regions on fracture
surface of a 5 mm diameter soda-lime-silica glass rod, tested in uniaxial tension. The mirror
region is roughly circular, surrounded by the narrow band of mist that gradually develops
into the hackle (after Johnson and Holloway, 1966).
274 6. Fracture: Basic Concepts and Experimental Techniques
two zones, then the fracture strength σf of the material, i.e., the stress at which
the crack starts to move (see also Chapters 7 and 8) is related to R through
√
σf R = a, R = Rmirror , Rmist , Rhackle , (19)
where a is a constant. Observe that Eq. (19) has √ the √same form as the Griffith
condition, Eq. (9) [if we rewrite Eq.√(9) as, σc c ∝ Y ], and therefore the
constant a is related to the quantity Y that appears in the Griffith condition.
Moreover, in view of Eq. (16), the constant a can also be related to the fracture
toughness Kc . Experiments have also indicated that if R0 is the radius of the
initial flaw at which the crack nucleates, then the radius Rmirror of the mirror
zone is related to R0 through, Rmirror /R0 10. Clearly, the circular boundaries
between the three zones will not develop if the crack cannot accelerate in all
directions with the same rate. The deviation from circularity depends partly on
the boundary conditions used in the test. For example, a material in a bending
experiment develops a stress distribution that is quite different from one that it
experiences in a uniaxial tension experiment. Moreover, the mechanism of crack
growth in amorphous materials is different from that of crystalline materials, so
that the shape of the boundaries between the mirror, mist and hackle zones also
depends on the material.
Before closing this section, let us point out that in the fracture literature one
often finds references to twist hackle and stress or velocity hackle. The former
refers to a rough surface that is generated by a Mode I/III fracture experiment,
whereas the latter is the result of a crack propagating at very high speeds or under
a large stress. The phrase mirror has also been used occasionally for describing the
initial stage of the development of a fracture surface, whereas careful examination
of the surface would reveal that it is too rough to be classified as the mirror zone.
To make the distinction between a mirror zone and a rougher region, one may
define the mirror region as the zone in which the average height of the roughness
is less that the wavelength of light.
Figure 6.8. Dynamic stress intensity factor Kd , the fracture velocity v, and the roughness
of the fracture surface versus the fracture length for a brittle epoxy resin. KI c is the critical
value of the stress intensity factor KI . The data are from Arakawa and Takahashi (1991)
(after Hull, 1999).
present on the surface, or are blunted by deformation, the mirror surface will not
develop. In addition, the presence of grain boundaries, multiple phases of the
material, and reinforcing particles force the crack paths into irregular shapes.
Systematic investigation of roughness of fracture surfaces and their scaling
properties were first undertaken by Mandelbrot et al. (1984), although Passoja
and Amborski (1978) and Chermant and Coster (1979) had already suggested
that fracture surface of metals may have fractal and scale-invariant properties. As
discussed in Chapter 1, if the width w of a rough surface follows the scaling law
(1.34), then the surface is a self-affine fractal with a fractal dimension Df which,
in d dimensions, is given by
Df = d − α, (20)
where α is the roughness exponent, which is usually the same as the Hurst ex-
ponent H introduced and discussed in Chapter 1, although, theoretically, the two
exponents can be different. Mandelbrot et al. (1984) studied fracture surface of
steel and concluded that the surface possessed fractal morphology. They estimated
the fractal dimension of the fracture surface of their material to be Df 1.28,
implying a roughness exponent α 0.72. If we assume that the roughness expo-
nent α is equivalent to the Hurst exponent H for the fractional Brownian motion
described in Section 1.4, a roughness exponent of 0.72 implies long-range positive
correlations on the fracture surface. Indeed, the profiles of such fracture surfaces
are very similar to fBm with a Hurst exponent H > 0.5 (see Figure 1.2). Since
the original work of Mandelbrot et al. (1984), many other measurements of frac-
tal and self-affine properties of fracture surface of a wide variety materials have
been reported. In particular, several experimental techniques have been used for
measuring and characterizing the roughness of fracture surfaces and estimating its
roughness exponent, which we now describe and discuss.
metals and their fracture toughness, such that Df decreased smoothly as the frac-
ture toughness increased. These issues and the progress up to 1988 were reviewed
by Williford (1988).
Mecholsky et al. (1988,1989) and Passoja (1988) studied fracture surfaces of
many solid materials, including several different aluminum and five glass ceramics,
all of which had distinct microstructures. They found that as the toughness of the
materials increases, so does also the roughness of the fracture surface. The fractal
dimension Df was found to be in the range 1.15 − 1.30, with an average of about
1.22, implying an average roughness exponent α 0.78. They also investigated
the relation between fracture energy and the geometry of fracture surface in many
different brittle materials and proposed the following equation
1
= Y ξ(Df − 1), (21)
2
where is the fracture energy, Y is an elastic modulus, and ξ is a characteristic
length scale of the material.
Dauskardt et al. (1990) undertook a systematic study of five samples of brittle
and ductile transgranular cleavage, intergranular fracture, microvoid coalescence,
quasi-cleavage, and intergranular microvoid coalescence in various steels. These
materials were fractured both at room temperature and also a very low tempera-
ture. They analyzed the measured length L of the surface versus the measuring step
1−D
length Ls which are related through, L ∼ Ls f . In many cases, a fractal dimen-
sion Df 1.2 was obtained, in agreement with the previous estimates discussed
earlier. However, in several other cases the relation between L and Ls was more
complex. Bouchaud et al. (1990) studied fracture of an aluminum alloy in 4 dif-
ferent heat treatment regimes. The fracture surface was elecro-coated with nickel,
then polished and digitized. The correlation function C(r), Eqs. (1.5)–(1.8), was
then constructed for the aluminum-nickel boundary for a large number of sam-
ples. Even though quite different mechanisms of fracture were dominant in these
materials, in all cases the roughness exponent was α 0.8.
Zhenyi et al. (1990) and Dickinson (1991) studied fracture surface of poly-
mers and ceramics, measuring both surface roughness and light emission signals.
Fractal dimensions of 1.2–1.3 were measured for the rough surfaces, resulting
in roughness exponents of about 0.7–0.8. The photon emission signals also had
fractal characteristics, and measurement of their fractal dimensions yielded values
between 1.24 to 1.42, implying roughness exponents in the range 0.6–0.75. Note
that, there appears to be a close relationship between the fractal dimensions of the
fracture surface and that of the emission signals. If the exact nature of this rela-
tionship can be identified, then photon emission signals may provide an accurate
probe of fracture surfaces and their morphology.
Fractures on carbon surfaces were analyzed by Miller and Reifenberger (1992),
who reported that α 0.75. Poon et al. (1992) studied fracture surface of natural
rock, such as sandstone, limestone, and carbonates. For each sample roughness
profiles of several thousand points were constructed, and for all cases studied
a roughness exponent of about 0.8 was obtained. Måløy et al. (1992) investi-
6.14. Roughness of Fracture Surfaces 279
gated fracture surfaces of six different brittle materials, ranging from Al-Si alloy
AA4253 to porcelain. The materials were notched and then fractured at the tem-
perature at which nitrogen becomes liquid. Many profiles of the rough fracture
surfaces were then obtained and analyzed. Two methods of analysis, including
the power-spectrum method described in Section 1.4.1, were used. The rough-
ness exponent was estimated to be α 0.87 ± 0.07 for all the six samples. Baran
et al. (1992) analyzed fracture surface of several brittle materials, including glass
and dental porcelain, and reported large roughness exponents, ranging from 0.65 to
0.93. Poirier et al. (1992) studied deformation of regular packings of equal parallel
cylinders. The local stress-strain characteristics, at the contact between the cylin-
ders, exhibited a softening part which localized the deformation. The deformation
band was rough with a roughness exponent α 0.73 ± 0.07.
An interesting method for studying fracture surface was developed by Imre et
al. (1992) who determined the fractal dimension of the surface electrochemically
by measuring the diffusion current, also called Cottrell current, at a gold replica
of the fractured metal electrode. (It is interesting to find research groups that are
rich enough to afford gold in their investigations, while others starve for research
funds!) The replicas were prepared by pressing gold wafers into the fractured
steel surfaces in a hydraulic press at high pressure. The gold surfaces were then
cleaned, and the gold electrodes were immersed in an aqueous electrolyte with a
calomel reference electrode. The potential was switched from 0 V to 650 mV for
a short period of time, and then was switched back to 0 V. According to Nyikos
and Pajkossy (1985) the current I (t) should scale with the time t as
I (t) ∼ t (α−2)/2 , (22)
so that simple measurements of I (t) versus t should yield α (and hence Df ).
Roughness exponents of about 0.8 were measured by this method.
Another interesting method for measuring roughness properties of a fracture sur-
face was developed by Friel and Pande (1993). In their method pairs of electron
micrograph images of fracture surface of titanium 6211 at two different inclina-
tion angles (30◦ and 36◦ ) were constructed using a scanning electron microscope
(SEM). The surfaces were fractured under tension. The SEM images were obtained
under various magnifications, ranging from 50 to 10,000. The surface fractal di-
mension was then estimated by measuring the surface area as a function of the
length scale (or measurement resolution), and was found to be about 2.22, implying
a roughness exponent α = 3 − 2.22 = 0.78. Schmittbuhl et al. (1993) measured
roughness exponent of several granitic faults and found α 0.85, close to the val-
ues obtained by others for various materials. E. Bouchaud et al. (1993b) analyzed
the statistics of fracture surfaces of polycrystalline intermediate compound Ni3Al.
Such fracture surfaces also contain secondary branches, as opposed to most of
the fracture surfaces discussed above which had no side branches. Despite this,
E. Bouchaud et al. (1993) could define a roughness exponent for fracture sur-
face of these materials, and their measurements indicated that α 0.8. Lemaire
et al. (1993) put a viscoelastic paste made of sand and resin between two plates
which were driven away from each other at a given velocity until the paste broke.
280 6. Fracture: Basic Concepts and Experimental Techniques
Five different velocities were used, and after fracture the hardened paste was
sliced parallel to the tensile direction. The fractal dimension of the profiles was
then determined by two methods, the standard box-counting method, and by the
power-spectrum methods, both of which were described in Chapter 1. A roughness
exponent α 0.88 ± 0.05 was measured which was independent of the velocity.
Daguier et al. (1995) studied the morphology of fractures in two different metal-
lic alloys. The fractures had been stopped during their propagation by pinning
microstructural obstacles to the surface. One of the alloys was the 8090-Al-Li
which is very anisotropic, for which the roughness exponent was found to be
α 0.6 ± 0.04. The other alloy was Super α2 Ti3Al with a 3D fatigue fracture for
which α 0.54 ± 0.03. Daguier et al. (1996) used atomic force microscopy and
SEM methods to study fracture surface of Ti3Al-based alloys. They found that at
large length scales, and over several decades in length scales, the roughness expo-
nent was α 0.8, whereas at much shorter length scales the roughness exponent
was close to 0.5. Daguier et al. (1997) also studied fracture surface of a silicate
glass as a function of the fracture velocity. At large length scales the roughness
exponent was α 0.78, whereas at smaller length scales α 0.5. The crossover
length scale ξco that separated the two scaling regimes was shown to be propor-
tional to the inverse of the fracture velocity. If hmax is the difference between the
maximum and minimum heights h within a given window on the surface, then the
two scaling regimes could be combined into a single scaling law
hmax ∼ r 0.5 (r/ξco ), (23)
where is a scaling function with the properties that (x) ∼ 1 as x → 0, and
(x) ∼ x 0.28 for x 1.
Thus, summarizing all the experimental data discussed so far, it appears that at
large enough length scales a roughness exponent α 0.8 represents a universal
value, regardless of the material or even the mechanism of fracture. The possi-
bility of universality of α was first pointed out by Bouchaud et al. (1990). We
should, however, point out that if a fracture surface is analyzed on relatively short
length scales, then the effective value of α may be smaller than 0.8. For example,
Mitchell and Bonnell (1990) analyzed fracture surface of fatigued polycrystalline
copper and reported that α 0.65, while for a single crystal silicon α 0.7 was
obtained. Metallic materials, the roughness exponents of which have been deter-
mined through scanning tunneling microscopy, usually operate in the nanometer
range and have α < 0.8. For example, Milman et al. (1993, 1994) reported a rough-
ness exponent of about 0.6 for fractured tungstene, and close to 0.5 for graphite.
Low cycle fatigue experiments on steel samples on micrometer scales yielded
a roughness exponent close to 0.6 (McAnulty et al., 1992). Low values of the
roughness exponents are interesting because they might be explained based on
models of minimum energy surfaces in disordered environments. Such concepts
were first discussed by Chudnovsky and Kunin (1987), Kardar (1990), Roux and
Francois (1991), and Ertas and Kardar (1992,1993,1994,1996). For example, Roux
and Francois (1991) argued that the path that is selected by a propagating fracture
should be such that the overall fracture energy is minimized. Their simulations un-
6.14. Roughness of Fracture Surfaces 281
der such a condition led to a roughness exponent in the range 0.4–0.5. The apparent
length-scale dependence of the roughness exponent α may also be explained in
another way based on the velocity of fracture propagation, and whether one is in
the regime of quasi-static or rapid fracture (Bouchaud and Navéos, 1995). This
distinction, its theoretical treatment, and the corresponding roughness exponents
will be described in Chapter 7, where we will also discuss the implication of the
self-affine structure of fracture surface for crack propagation.
Now that there is little doubt that fracture surface of a wide variety of materials
is rough with well-defined characteristics, let us briefly describe how such sur-
faces are studied experimentally. This subject has been discussed in detail by Hull
(1999), and what follows is a summary of his discussion. Roughness is typically
characterized by measuring the height h of the roughness profile. A “primitive”
method would be based on using a raster scan of parallel traverses across the sur-
face using a stylus which traverses parallel to the x-axis—the axis that is parallel
to the mean position of the roughness profile—and measures the height. The stylus
is typically a fine, diamond-tipped needle which is in contact with the surface by a
small external load. The height of the needle is measured using a transducer. The
disadvantage of this method is that the stylus may damage the surface, and hence
create traces that do not belong to the original fracture surface.
Atomic force microscope can also be used which has a highly fine silicon nitride
stylus with a tip radius of about 20–30 nm. The probe is held at a fixed position
from the base of the rough surface, and the surface itself is moved parallel to this
base. The height of the probe is measured from the reflection of light from mirror
on the stylus beam. A powerful feature of this method is that it can determine
roughness parameters on specific sections of the roughness profile.
In a modern version of the stylus technique, the mechanical stylus is replaced
by a fine laser beam that is held at a constant distant from a references surface.
The size of the spot is typically 1 µm in diameter, and the rough surface traverses
under the beam light. The surface shape is then determined from the change in the
length of the light’s path that is reflected from the surface.
detailed discussions), with the scale of their roughness being equal to at least the
scale of the microstructure. We will come back to this mechanism in Chapter 7,
where we discuss the relation between micro-cracking and dynamics of fracture
propagation.
the effective properties of the material, such as their elastic moduli and fracture
toughness, but also an understanding of such micro-deformation processes as slip
that usually precedes and accompanies fracture in crystalline materials. The de-
gree of symmetry that the crystalline material exhibits also plays an important role,
because the strength of the anisotropy of micro-deformation processes depends on
such symmetries. The most important effect of anisotropy is that cleavage may
occur parallel to planes in a crystal that are not normal to the maximum tensile
stress. This is particularly true in crystalline materials that exhibit a low degree
of symmetry, such as mica in which cleavage is only in a single set of planes. In
addition, temperature and strain rates also play important roles by influencing the
mobility of dislocations.
The low surface energy of crystallographic planes, which in turn depends on
the strength of the interatomic bonds, is the main cause of cleavage in crystals. If
cleavage occurs along a single plane, it would produce a featureless surface. How-
ever, often one observes well-defined and crystallographically oriented features
on the fracture surface of a crystalline material. These features are usually caused
by the generation and presence of dislocations that interact with the propagating
fracture. In metals with body-centered cubic symmetry, such as chromium, tung-
sten, and iron, the main cleavage occurs on {001} planes, of which there are three,
(001), (010), and (100). If a cyrstal is tested in an arbitrary direction, the {001}
plane with the largest tensile stress normal to the plane is the most likely place
for cleavage. If a crystal is tested in tension parallel to [011], the (001) and (010)
planes have the same resolved normal tensile stress. In this orientation the stress
on the (011) plane is much greater than on the {100} planes. Thus, fracture may
occur either on an (011) plane, or along the two equally stressed {001} planes.
On the other hand, crystals with the zinc-blende structure, such as gallium ar-
senide, can be described as a cubic unit cell that consists of two interpenetrating
FCC lattices of the two elements (Ga and As). The center of one lattice is at the po-
sition (1/4,1/4,1/4) of the other. These materials are of great industrial importance
because of their use in producing semi-conductors. They cleave on {001} planes,
of which there are three equivalent pairs of orthogonal planes. Slip is restricted to
{111} planes. Such materials usually exhibit strong brittleness.
If polished (001) faces of GaAs crystals are coated with an epitaxial layer of
GaAs that contains a small amount of carbon, tensile stresses are generated in the
surface layers. These stresses then lead to the formation of very fine, atomically
sharp surface cracks (see, for example, Murray et al., 1996). The cracks form on
two orthogonal {011} planes that intersect the (001) surface at right angles, remain
sharp, and grow at very low stresses. The fracture surface is mirror smooth and flat.
However, if GaAs crystals are tested in complex loading conditions, the fracture
surface becomes very rough.
Layered materials usually have very strong bonding within the layers and weak
bonding between the layers. An example is muscovite mica that consists of an
ordered stack of double layers, about 2 nm thick, of strongly bonded planar arrays of
silica tetrahedra held together by Coulomb attraction caused by the potassium ions
between the layers. In such materials cleavage occurs between the weakly-bonded
284 6. Fracture: Basic Concepts and Experimental Techniques
layers, and may also occur through the center of the double layer. Deformation is
restricted to the sliding of layers over each other. The reader should consult Hull
(1999) for extensive discussions of other crystalline materials.
So far we have discussed the cleavage of single crystals. In practice, cleavage
of polycrystalline materials, such as ceramics and rock, is also very important.
Let us briefly discuss these phenomena. We assume that the bonding between the
crystals is very strong, and that the grain boundary interface does not experience
failure.
In polycrystalline materials, each grain is surrounded by many other grains of
different orientations. Therefore, such materials fracture by successive nucleation
and propagation of several cleavage cracks across the boundaries between neigh-
boring crystals. There is a change in the orientation at the grain boundary. If the
angle between the neighboring grains is small, the cleavage crack in a crystal can
propagate across the boundary between the neighboring crystals, in which case
the cleavage plane is tilted and twisted. However, if the orientations of the crystal
grains are very different, the propagation of cleavage from one crystal to another
depends on the relative orientation of the cleavage planes in the crystals.
Consider, for example, two adjacent grains with a common boundary between
them, and suppose that a crack in one of the grains reaches the boundary. Then, it
may stop there with no further crack propagation. Alternatively, the crack may stop
at the boundary, but the high stress at its tip may help nucleate another crack in the
adjacent grain with a different orientation. The two cracks have a common point at
the boundary. The third possibility is having a cleavage plane in the second grain
that is tilted relative to the cleavage plane in the first grain, in which case the crack
propagates continuously across the boundary. Therefore, fracture propagation in
polycrystalline materials depends critically on the distribution of their grains or
single crystals. Even if an array of grains is distributed randomly, the local direction
of crack propagation depends on the relative orientations of the grains at the crack
tip. On the scale of the single crystal size, the main crack path is not straight. It
is also possible that local regions of the crack “tunnel” ahead of the main crack
front because of the existence of a path of favorably oriented single crystals in
the region. If a polycrystalline material contains preferred orientations, then crack
growth in it is easier in some directions than others.
Figure 6.9. A typical stress-strain diagram for polymers. The top curve corresponds to
brittle behavior, while the middle curve leads to ductile behavior. In the lowest curve, strain
hardening leads to brittle behavior.
286 6. Fracture: Basic Concepts and Experimental Techniques
logarithmically with the strain rate, and slowly decreases with increasing temper-
ature, eventually vanishing at Tg .
Between the elastic limit and the yield point, many polymers that are under ten-
sion exhibit a series of crazes that are normal to the tensile stress. Both amorphous
and crystalline polymers generate crazes with the same features. In particular, it
is easy to see crazes in amorphous polymers as they strongly scatter visible light.
The inside of a polymer craze is typically filled with polymer fibrils, as a result of
which the effective moduli of the material after crazing is only slightly smaller than
before, implying that the onset of crazing cannot be detected on the stress-strain
diagram.
Under tension or compression, polymeric materials can also develop shear
bands, i.e., zones of highly localized shear. The bands are diffuse at high tem-
peratures or low strain rates, but are localized at lower temperatures or higher
strain rates. If the diffuse bands are further deformed, it will lead to ductile frac-
ture, whereas deformation of localized shear bands leads to brittle fracture. If two
shear bands intersect, it usually leads to a craze. The stress at the craze tip can also
lead to the formation of shear bands.
6.16.2 Ceramics
The British Ceramic Society defines ceramic materials as, “All solid manufac-
tured materials or products that are chemically inorganic, except for metals and
their alloys, and which are usually rendered serviceable through high temperature
processing.” Ceramic materials include borides, carbides, halides, nitrides, oxides,
and cermets, which are ceramic metals. They usually have a crystalline structure,
but can also be found in amorphous form. The interatomic bonds in ceramics may
be ionic, covalent, metallic, and van der Waals. It is clear how the first two types of
bonds may form in ceramics. Metal transition carbides have bonds which have a
metallic characteristic in that, valence electrons are freely shared by all the atoms
in the structure.
Relative to metals, ceramics have large elastic moduli, ranging from 70 to 400
GPa. The moduli decrease very slowly with increasing temperature. They also
have a large cohesive strength which is due to the fact that their interatomic bonds
require high energies to be broken. However, as discussed earlier in this chapter,
the presence of defects, which results in stress concentration, reduces the actual
strength of these materials. In fact, the fracture strength σf of ceramics is very
sensitive to the presence of defects, the porosity, the shape and size of the grains, as
wells as the pore-crack combination. Most importantly, σf depends on the size of
the defects, for which there is a critical size that, at a given stress, leads to fracture.
For ceramics this size can be as large as a single crystal. The Weibull distribution
[see Eq. (5.37); see also Chapter 8] usually describes well the statistical distribu-
tion of the fracture strength of ceramics. Moreover, if the defects are uniformly
distributed in the material, the probability of having the critical condition in the
material for fracture is relatively large. Experiments have indicated that in many
ceramics, especially those that have a secondary phase, the crack velocity v is
related to the stress intensity factor KI by, v ∼ KIn , where n is a constant.
6.16. Fracture Properties of Materials 287
Experiments by Buresch et al. (1983) and others have also shown that the fracture
strength of certain ceramics depends on the critical value σn of the notch fracture
stress, and also on the size of the cohesive zone (see above). The cohesive zone in
ceramics is somewhat similar to the plastic zone in metals in that, the microcracked
zone in the immediate vicinity of a crack tip causes the nonlinear behavior of
ceramics. In this zone, there is a constant stress σn for breaking either the grain
boundaries or the crystal themselves, which depends on the cohesive stress σc (see
above). If the average stress in the cohesive zone reaches σn , instability occurs in
the material and the main crack propagates.
The behavior of the fracture strength σf of ceramics with variations in the tem-
perature can be divided into two groups. In one group, σf decreases monotonically
with increasing temperature. Nitrides typically exhibit this behavior. In the second
group, the fracture strength either stays constant with increasing temperature, or
first experiences a small increase and then decreases. Ceramics that do not have a
secondary phase at their grain boundaries exhibit this behavior.
6.16.3 Metals
Most metals have simple crystalline structures in the form of BCC or FCC lattices
or a hexagonal close-packed (HCP). At the atomic scale the interatomic bonds
break either along crystallographic plane in Mode I fracture (i.e., in a direction
normal to the plane), or in Modes II and III fracture (i.e., in a direction parallel to
that plane), which is also the mechanism for cleavage fracture already described
above. Alternatively, metals fracture at high temperatures by coalescence of cavi-
ties. Single crystals of a HCP metal (for example, zinc) can slip on a single plane
until the two parts completely separate. Usually, however, multiple slip occurs
in single crystals which generates a neck in the material which is under tension.
These necks usually initiate at inclusions which do not deform in the same way
as the metallic matrix. In polycrystalline metals, necking occurs in a more diffuse
fashion, but can lead to the complete separation of the two halves of the material
when the neck’s cross section vanishes. This mechanism is, however, rare. In most
cases, the material breaks much sooner by developing a crack in the middle of
the neck which is perpendicular to the tensile axis, which at the end tilts to a 45◦
orientation. This crack is the result of coalescence of vacancies which grow due to
plastic deformation and elongate in the direction of the maximum principal strain.
This mechanism of fracture in metals is ductile because it involves large local
slip deformation. It also often corresponds to a large macroscopic plastic de-
formation. However, coalescence of the vacancies does not always need large
macroscopic deformation, such as when the volume fraction of the inclusions in a
metal-matrix composite is large. Hydrostatic pressure prevents the growth of the
cavities, whereas a tensile positive hydrostatic stress increases it, and thus reduces
greatly the fracture strain.
Another mechanism of fracture of metals is intergranular cracking, which hap-
pens when the grain boundaries are weaker than their interior. The weakness is
caused by impurities that have accumulated at the grain boundaries. In such a situ-
ation, the cracks preferentially follow the grain boundaries, leading to intergranular
288 6. Fracture: Basic Concepts and Experimental Techniques
(3) coalescence of the cavities due to plastic deformation and formation of a crack,
and
(4) propagation of the crack.
Inside the metallic matrix the failure is ductile, but it appears brittle at the
macroscopic length scales.
An important factor is the aspect ratio of the fibers, i.e., the ratio of their lengths
and diameters. For example, fibers reinforce a material better than spherical parti-
cles. The larger the aspect ratio, the higher is the fracture strength of the composite.
However, if the fibers become too long, they will no longer influence the strength
of the material. The properties of the interface between the fibers and the matrix
also have a very strong effect on the strength of the composite. If the interface is
stronger, the composite material will have a lower ductility and a higher fracture
strength. During production of the composite, internal stresses may be produced
by mismatch between the thermal expansion coefficients of the matrix and the
fibers. Thus, when the temperature of the system is reduced, residual stresses are
produced in the composite which, however, disappear at high enough tempera-
tures. In addition, one may have chemical reaction at the interface. If, for example,
the fibers are oxidized, the fracture strength of the composite will reduce.
Summary
The aim of this chapter was to define the basic concepts of fracture mechanics, and
describe and discuss the basic phenomena that occur during fracture of materials.
These concepts will be utilized in the next few chapters where we describe and
discuss modeling of brittle fracture of heterogeneous materials and its transition
to ductile behavior. We also described the experimental techniques that are used
for measuring important characteristics of fracture of materials, such as the speed
of crack propagation, and measurement and analysis of roughness of a fracture
surface.
7
Brittle Fracture: The Continuum
Approach
7.0 Introduction
As discussed in the last chapter, fracture of brittle amorphous materials is a difficult
problem, because the way a large piece of a material breaks is closely related to
details of cohesion at microscopic length scales. For this reason alone, description
of brittle fracture of materials has been plagued by conceptual puzzles. What made
matters worse for a long time was the fact that many past experiments seemed to
contradict the most firmly-established theoretical results. However, considerable
progress has been made over the past decade, and one main aim of this chapter
is to demonstrate that the theory and experiments fit within a consistent picture.
This has become possible by the realization that dynamic instabilities of the tip of
a fracture play a critical role in determining the fracture behavior of amorphous
materials. To accomplish this goal, we follow our by-now-familiar path, namely,
we first describe and summarize the central results of continuum theories of linear
elastic dynamic fracture mechanics which provides an elegant and powerful de-
scription of fracture propagation. However, the continuum theory is unable to make
quantitative predictions without additional information that must be provided by
experiments, or be supplied by other types of theories. We already discussed in
the last chapter some of the most important experimental observations and data,
and the techniques that were used for obtaining them. These experiments teach us
that when the flux of energy to a fracture tip exceeds a critical value, the fracture
becomes unstable and hence propagates in an increasingly complex manner. As a
result, the moving crack cannot travel as quickly as the linear continuum theory
predicts or assumes, the fracture surface becomes rough and begins to branch out
and radiate sound, and the energy cost for the motion of the crack increases signif-
icantly. These observations are completely consistent with the continuum theory,
but cannot be described by it. Therefore, to complete the emerging theoretical
picture and the fundamental understanding of this phenomenon, we continue this
chapter with an account of theoretical and numerical work of the past decade or
so that attempts to explain the dynamic instabilities in fracture propagation. As
discussed in the last chapter, our current experimental understanding of instabili-
ties in fracture tip in brittle amorphous materials is fairly detailed. We also have
a rather detailed theoretical understanding of these instabilities in crystals which
reproduces many qualitative features of the experiments. Recent numerical work
7.0. Introduction 291
goal of this chapter is to explain how these puzzles have arisen, and how to recast
them in new terms and explain them. We do not intend to provide a complete
review of fracture mechanics as it will require a book by itself. Instead, we focus
on brittle materials. Ductility and dynamic elasto-plastic fracture, which is a well-
developed field, have been described well by others (see, for example, Freund,
1990; Chan, 1997). Therefore, we will discuss only the brittle-to-ductile transition.
The emphasize in this chapter is first to describe and summarize the most important
predictions of the conventional continuum fracture mechanics, and then answer
some fundamental and interesting questions that this type of models do not ask or,
if they do, cannot answer. To write a significant portion of this chapter, we relied
heavily on the excellent review of this subject by Fineberg and Marder (1999).
Some of the developments that we discuss had been described in an earlier article
by the author (Sahimi, 1998), and thus have also been utilized in this chapter.
As is well known, interatomic forces vary greatly between different elements and
molecules (see Chapter 9), but they typically attain their maximum value when
the distance between atoms increases by about 20% of its original value. The force
needed to stretch a solid material slightly is, F = Y Sδ/L, where Y is the Young’s
modulus, L is its initial length, and δ is the amount (in length) that the material
has been stretched. Therefore, the force per unit area needed to reach the breaking
point is about, σc = F /S = Y δ/L Y /5, where we have used δ/L = 0.2. We
list in Table 7.1 values of Y for several materials, the theoretical strength σc , and
its comparison with the experimental data. As this table indicates, the theoretical
estimate of σc is in error by orders of magnitude. The scaling estimate of hc greatly
underestimates the practical resistance of solid materials to fracturing, while the
estimate of σc too large. What is the problem? The only way to discuss the correct
orders of magnitude is to account for the actual dynamical mode by which brittle
materials fail mechanically, which is by propagation of a fracture.
As described in Chapter 6, and shown later in this chapter, the presence of a
fracture in an otherwise perfect material results in a stress singularity at the frac-
ture tip. If the fracture tip is atomically sharp, a single fracture which is a few
microns long suffices for explaining the large discrepancies between the theoret-
ical and experimental material strengths that are shown in Table 7.1. The stress
singularity that develops at the tip of a fracture focuses the energy that is stored
in the surrounding material and uses it efficiently for breaking one atomic bond
after another. Thus, continuous fracture propagation provides an efficient way of
overcoming the energy barrier between two equilibrium states of the system that
have different amounts of mechanical energy. We now turn to a scaling analysis
of dynamic fracture (Fineberg and Marder, 1999).
analysis which, despite being wrong in many of its details, clarifies the basic
physical processes. It consists of writing down an energy balance for the motion
of a fracture. Consider a fracture of length l(t) growing at time t at rate v(t) in a
very large plate to which a stress σ∞ is applied at its far boundaries; see Figure
7.1. As the fracture extends, its faces separate, causing the plate to relax within a
circular region centered at the middle of the crack with a diameter which is of the
order of l. The kinetic energy Hk involved in moving a piece of material of this
size is 12 mv 2 , where m is the total mass, and v is a characteristic velocity. Since
the mass of the moving material is proportional to l 2 , the kinetic energy should be
given by
Hk (l, v) = ck l 2 v 2 , (1)
where ck is a constant. The moving portion of the material is also where elastic
potential energy is being released as the crack propagates. This stress release results
in a gain in the potential energy which is given by
Hp (l) = −cp l 2 , (2)
where cp is another constant. Equations (1) and (2) are correct if the crack moves
slowly, but they fail even qualitatively if the fracture velocity approaches the speed
of sound, in which case Hk and Hp both diverge. Their divergence will be demon-
strated below, but let us assume for now that it is true. The final piece of the energy
balance is the contribution of creation of new fracture surfaces. This contribution
is l, where is the fracture energy that, as described in Chapter 6 (see Section
6.7), accounts for the minimum energy needed to break the atomic bonds and any
other dissipative processes that the material may need in order for the fracture
to propagate, and is often orders of magnitude greater than the thermodynamic
surface energy. Therefore, the total energy H of the system containing a fracture
7.1. Scaling Analysis 295
is given by
H(l, v) = ck l 2 v 2 + Hqs (l), (3)
where Hqs is the quasi-static part of the total energy given by
Hqs (l) = −cp l 2 + l. (4)
If a crack moves forward slowly, its kinetic energy will be negligible, and therefore
only Hqs will be important. For small fractures, l, the linear cost of fracture
energy, is always greater than the quadratic gain of the potential energy, Hp = cp l 2 .
In fact, such fractures would heal (move backward) if such irreversible processes
as oxidation of the crack surface did not prevent them from healing. The fact that
the fracture grows is due to additional irreversible processes, such as chemical
attack on the crack tip (see Chapter 6), or vibration and other irregular mechanical
stresses. Eventually, at a critical length lc , the energy gained by relieving elastic
stresses in the material exceeds the cost of creating new fracture surfaces, in which
case the crack is able to extend spontaneously. Clearly, at lc , the energy functional
Hqs (l) has a quadratic maximum. The Griffith criterion (Griffith, 1920; see Chapter
6 and also below) for the onset of fracture is that fracture occurs when the potential
energy released per unit crack extension equals the fracture energy . Thus, fracture
in this system occurs at a critical crack length lc such that, dHqs /dl = 0 at l = lc .
Using Eq. (4) we find that,
lc = , (5)
2cp
so that
Hqs (l) = Hqs (lc ) − cp (l − lc )2 . (6)
The most important issue in engineering fracture mechanics is calculating lc ,
given such information as the external stresses which, in the present case, is rep-
resented by the constant cp . Dynamic fracture begins in the next instant, and since
it is very rapid, the energy H of the system is conserved, remaining at Hqs (lc ).
Thus, from Eqs. (3) and (6) we obtain
$ % $ %
cp lc lc
v(t) = 1− = vm 1 − , (7)
ck l l
which predicts that fracture propagation will accelerate until it approaches the
maximum speed vm . Equation (7), and more generally the above scaling analysis,
cannot by themselves predict vm , but Stroh (1957) argued correctly that vm should
be the Rayleigh wave speed cR , although his suggestion was implicitly contained
in the earlier calculations of Yoffe (1951) (see below).
In this system, one needs only to know the length lc at which a fracture begins
to propagate in order to predict all the ensuing dynamics. As we discuss later
in this chapter, Eq. (7) is actually very close to anticipating the results of a far
more sophisticated analysis, which is surprising since the Eqs. (1), (3) and (4)
for the kinetic and potential energy are in fact incorrect because they actually
296 7. Brittle Fracture: The Continuum Approach
diverge as the speed of fracture propagation approaches the Rayleigh wave speed
cR . However, the success of Eq. (7) is due to the fact that it involves the ratio
Hp /Hk . Since the divergence of the kinetic and potential energy are according to
exactly the same forms, the errors involved in their estimation cancel each other
out. We now attempt to review and discuss the background, basic formalism and
underlying assumptions that form the basis for continuum fracture mechanics.
the size of the cohesive zone depends on the material, ranging from nanometers
in glass to microns in brittle polymers.
(2) The universal elastic region, which is the region outside the cohesive zone
for which the response of the material can be described by linear continuum
mechanics. Outside the cohesive zone, but in the vicinity of the fracture tip,
the stress and strain fields take on universal singular forms which depend only
on the symmetry of the externally applied loads. In two dimensions (2D) the
singular fields surrounding the cohesive zone are completely described by
three constants which are the stress intensity factors introduced and discussed
in Section 6.8 (see also below). They incorporate all the information regarding
the loading of the material.
(3) The outer elastic region far from the crack tip in which stresses and strains are
described by linear elasticity. Details of the solution to the stress field in this
region of materials depend only on the locations and strengths of the loads,
and the shape of the material. For some special cases analytical solutions have
been derived. Deriving such solutions is made possible by the fact that, so far
as linear elasticity is concerned, viewed on macroscopic scales, the cohesive
zone can be represented by just a point at the fracture tip, while the fracture
itself is equivalent to a branch cut. In general, however, one must resort to
numerical simulations and solutions.
The dissipative processes within the cohesive zone determine the fracture energy
. If no dissipative processes other than the direct breaking of the atomic bonds
take place, then is a constant which depends on the bond energy. In general
though, is a complex function of both the fracture velocity and history, and
differs by orders of magnitude from the surface energy—the amount of energy
required to sever a unit area of atomic bonds. No general first principle description
of the cohesive zone exists, although numerous models have been proposed (see,
for example, Lawn, 1993).
the elastic fields throughout the material. However, the stress that locally drives the
fracture is one which is present at its tip. Thus, Kβ determines entirely the behavior
of a fracture, and much of the study of fracture processes is aimed toward either
calculating or measuring this quantity. The universal form of the stress intensity
factor allows a complete description of the behavior of the tip of a fracture where
one needs only carry out the analysis of a given problem within the universal elastic
region (see below). For arbitrary loading configurations, the stress field around the
fracture tip is given by three stress intensity factors Kβ which lead to a stress field
that is a linear combination of the pure Modes:
3
Kβ β
σij = √ fij (v, θ ). (9)
β=1
2π r
As mentioned above, the stress intensity factors are related to the flow of energy
into the fracture tip. Since a fracture may be viewed as a means of dissipating built-
up energy in a material, the amount of energy flowing into its tip must influence
its behavior. Irwin (1956) showed that the stress intensity factor is related to the
energy release rate H, defined as the amount of energy flowing into the crack tip
per unit fracture surface formed. The relation between the two quantities is given
by
3
1 − νp2
H= Aβ (v)Kβ2 , (10)
Y
β=1
where νp is the Poisson’s ratio of the material, and the three functions Aβ (v)
depend only on the fracture velocity v. Equation (10) is accurate when the stress
field near the tip of a fracture can be accurately described by Eq. (8), which is the
case as the dimensions of the sample increase.
stress elasticity fail (Nakamura and Parks, 1988). If the thickness of the plate along
the z-direction is denoted by d, then at distances from the fracture tip that are much
larger than d all fields are described by equations of plane stress. At distances from
the fracture tip that are much less than d, and away from the x − y surfaces of the
plate, the fields solve the equations of plane strain.
Equations (41)–(43) represent the general solutions in which the functions φ and
ψ must match the boundary conditions that are specified. Since one wishes to find
the potentials from given stresses at the boundaries, must diverge as 1/v, and
the right-hand sides of Eqs. (41)–(43) turn into the derivative of with respect
to α, implying that the static theory has a different structure than the dynamical
theory which is in fact more straightforward.
Let us now derive, as an application of Eqs. (37)–(43), the expressions for the
stresses around the tip of a fracture moving under Mode I loading. We assume that
the fracture lies along the negative x-axis (terminating at x = 0) and propagates
forward. The only assumption is that the problem is symmetric under reflection
about the x-axis. As discussed above (and also in Chapter 6), in the static case, the
stress fields have a square root singularity at the crack tip. We assume the same
to be true in the dynamic case (which can be verified in all cases for which the
304 7. Brittle Fracture: The Continuum Approach
expressions have been derived). Therefore, we assume that near the fracture tip
(Fineberg and Marder, 1999)
φ(z) ∼ (br + ibi )z−1/2 , (44)
−1/2
ψ(z) ∼ (dr + idi )z . (45)
Since we are considering Mode I fracture, then by symmetry the displacements
satisfy
ux (−y) = ux (y), uy (−y) = −u(y). (46)
If we substitute Eqs. (44) and (45) into (39) and (40) and use Eq. (46), we find that
bi = dr = 0, and therefore
(z) ∼ br z−1/2 , (47)
−1/2
(z) ∼ idi z . (48)
Observe that the square roots in Eqs. (44) and (45) must be interpreted as having
their cuts along the negative x-axis, where the fracture is located. Since on the
crack surface the stresses are relaxed, σxy and σyy vanish there. If we substitute
Eqs. (47) and (48) into Eqs. (41)–(43), we find that the condition for σyy is satisfied
identically for x < 0, y = 0, and that at y = 0
√ √
σxy = iµ 2αbr − (β 2 + 1)di (1/ x − 1/ x), (49)
and therefore
di 2α
= 2 , (50)
br β +1
which, when used in Eqs. (41)–(43), (47) and (48), yields
$ %
KI 1 1 1 1
σxx = √ (β + 1)(1 + 2α − β ) √ + √
2 2 2
− 4αβ √ + ! ,
2π D zα z̄α zβ z̄β
(51)
$ %
KI 1 1 1 1
σyy = √ 4αβ √ +! − (1 + β 2 )2
√ +√ , (52)
2 2π D zβ z̄β zα z̄α
2iαKI 1 1 1 1
σxy = √ (β + 1) √ − √ − √ + !
2
, (53)
2 2π D zα z̄α zβ z̄β
with
D = 4αβ − (1 + β 2 )2 . (54)
Note that the Rayleigh wave speed is in fact the root of D = 0, when Eqs. (35)
and (36) are used in (54). The most important physical feature of Eqs. (51)–(53)
is the overall scale of the stress singularity, which is characterized by the Mode I
stress intensity factor which, at y = 0, is given by
√
KI = lim 2π xσyy , (55)
x→0+
7.3. Linear Continuum Theory of Elasticity 305
σ?
σ OO
σ?
Figure 7.2. Behind its tip, a fracture is pulled apart by two stresses (after Fineberg and
Marder, 1999).
which, as will be shown below, is directly related to energy flux into a fracture
tip. Moreover, Eqs. (51)–(53) contain information about the angular structure of
the stress fields which can be used in both theoretical and experimental analyses.
Theoretically, one can use these equations for predicting the direction of fracture
motion, and the conditions under which a fracture branches out. Experimentally,
one can utilize these equations for assessing the accuracy of the predictions of
continuum fracture mechanics, and for obtaining measurements of the stress fields
surrounding rapidly-propagating fractures; we will discuss these matters later in
this chapter. It is important to recognize, as pointed out by Freund (1990), that
although Eqs. (51)–(53) were derived for fractures moving at a constant speed, the
same equations are also true for those that, during propagation, accelerate and/or
decelerate, so long as the derivative dv/dt is small during the time needed for
sound to travel across the region of the universal elastic singularity.
We now suppose that a fracture is loaded by two stresses, located a distance l0
behind its tip, moving with it in steady state at velocity v, and of strength −σc (see
Figure 7.2) such that
lim σyy (x, y) = −σc δ(x + l0 ), x < 0. (56)
y→0+
If the fracture tip is at the origin, the stress and displacement fields are continuous
and differentiable everywhere, except along a branch cut starting at the origin and
running backwards along the negative x-axis. If we define ± (x) and ± (x) by
± (x) ≡ lim (x ± iy), ± (x) = ± (x ± iy), (57)
y→0+
then because of the branch cut, for x < 0, + (x) = −− (x). As shown above,
for Mode I loading, σxy = 0 for y → 0+ and ∀x. Therefore, from Eq. (43) we
obtain
¯ − ) = (β 2 + 1)(+ +
2iα(+ − ¯ − ), (58)
306 7. Brittle Fracture: The Continuum Approach
crack is equal to , the energy required for creating new fracture surface. Griffith’s
idea, which is the final assumption of continuum fracture mechanics, states that
the dynamics of a fracture tip depends only on the total energy flux H per unit area
into the cohesive zone, and that all the details about the spatial structure of the
stress fields are irrelevant. The energy H creates new fracture surfaces, and is also
dissipated near the fracture tip. In general, the fracture velocity v is a function of
H. It is common to use (v) for representing the energy consumed by a fracture
in the cohesive zone, in which case the equation of motion for a fracture is
H = (v). (67)
The central question of interest to continuum fracture mechanics is the conditions
under which a static fracture begins to move. For this to happen, a critical fracture
energy Hc , the minimum energy per unit area needed for a fracture to propagate
forward, irrespective of its velocity, is needed. The standard assumption is that the
velocity consuming the minimum energy is very small, although this assumption is
not necessarily correct. Equivalently [see Eq. (10) and Chapter 6], one may define
a critical stress intensity factor KI c at which the fracture first begins to propagate.
We now derive this equivalence, following Fineberg and Marder (1999).
In what follows, we adopt the summation convention for repeated indices.
Energy flux is found from the time derivative of the total energy:
$ %
d d 1 1 ∂uα
(Hk + Hp ) = ρ u̇α u̇α + σαβ dxdy, (68)
dt dt 2 2 ∂xβ
where Hk and Hp are, respectively, the total kinetic and potential energies within
the entire system, and u̇α = duα /dt. Since the spatial integral in Eq. (68) is taken
over a region which is static in the laboratory frame (i.e., dx/dt = dy/dt = 0),
we have
$ %
d ∂ u̇α
(Hk + Hp ) = ρ üα u̇α + σαβ dxdy, (69)
dt ∂xβ
where the symmetry of the stress tensor under interchange of indices has been used
for the last term. Use of the equation of motion, ρ üα = ∂σαβ /∂xβ , in Eq. (69),
yields,
$ %
∂ ∂ u˙α ∂
σαβ u̇α + σαβ dxdy = σαβ u̇α dxdy
∂xβ ∂xβ ∂xβ
= u˙α σαβ nβ dS, (70)
∂S
where ∂S is the surface boundary of the system, and n is an outward unit nor-
mal with components nβ . Equation (70) is a statement of the fact that energy is
transported by a flux vector j with components that are given by
jα = σαβ u̇β . (71)
As mentioned in Chapter 5 [see Section 5.1.1 and Eq. (5.13)], the total energy
flux J per unit time into the fracture tip is called the J -integral (see Cotterell
308 7. Brittle Fracture: The Continuum Approach
Figure 7.3. Dotted lines show the most convenient contour for integrating the energy flux
and calculating the energy that flows to a fracture tip. The contour runs below the fracture,
closes at infinity, and comes back just above the contour (after Fineberg and Marder, 1999).
and Atkins, 1996, for a discussion of the use of the J -integral to ductile fracture).
A convenient contour for the integration is shown in Figure 7.3. If, for a crack
loaded in pure Mode I, we use the asymptotic forms, Eq. (52) for σyy and the
corresponding expression for uy , we find that J is given by
α v(1 − β 2 )
J = K 2, (72)
2µ 4αβ − (1 + β 2 )2 I
where KI is the stress intensity factor defined by Eq. (55), with the subscript I
emphasizing that the result is specific to Mode I fracture. Thus, the energy release
rate H in the case of pure Mode I is
J α 1 − β2 1 − νp2
H= = K 2
≡ AI (v)KI2 . (73)
v 2µ 4αβ − (1 + β 2 )2 I Y
The corresponding result for pure Mode II fracture is
β 1 − β2 1 − νp2
H= K 2
≡ AI I (v)KI2I , (74)
2µ 4αβ − (1 + β 2 )2 I I Y
while for Mode III fracture one has
v
H= K2 . (75)
2αµ I I I
In the limit v → 0, each of the functions Aα (v) → 1 (α =I, II and III) and, for
example, Eq. (73) simplifies to
1 − νp2
H= KI2 . (76)
Y
In the general case of mixed mode fracture, Eq. (10) should be used.
The functions Aα (v) are universal in the sense that they are independent of
most details of the material’s loading or geometric configuration. Assuming that
there is no energy sink in the system other than the one at the tip of the fracture,
7.5. The Equation of Motion for a Fracture in an Infinite Plate 309
Eqs. (73)–(75) relate the total flux of energy from the entire elastic material to the tip
which, when it is set to equal to the energy dissipated in the cohesive zone, yields
an equation of motion for the fracture. Note that, in order to derive Eqs. (73)–
(75), we have tacitly assumed that, given near-field descriptions of stress and
displacement fields [Eqs. (50)–(53)], Eqs. (37) and (38) are valid. If, for example,
the cohesive zone is of the order of 1 mm in a piece of a solid with dimensions
that are a few centimeters, the value of the stress field on the contour ∂S used in
Eq. (70) will not be approximated well by the asymptotic forms of the stress and
displacement fields, invalidating Eqs. (73)–(75). Since an energy balance provides
no information about a fracture’s path, we have assumed that the fracture travels
along a straight line (see below). Although the rules for determining paths of slowly
propagating fractures are known, they are not known for rapidly moving fractures.
is assumed to be known. However, v(t) must be less than the relevant sound
speeds at all times.
(3) The external stresses σe are permitted only along the fracture, but are allowed
arbitrary time and space dependence. This can be realized by placing wedges
between the faces of the cracks in order to load them.
We derive the corresponding expressions for fracture Modes I, II and III. In the
calculations that follow u, σ and c denote the displacement, stress, and a sound
speed in each case, as listed in Table 7.2. By symmetry, u(x, t) = 0 for all x > l(t).
Due to the one-to-one relation between K and the energy flux H, we compute the
latter as a function of l(t) and v(t), and as a functional of the external load σe (x, t).
We look for a Green function G operating on the displacement field u defined by
the following convolution integral,
G∗u≡ G(x − x , t − t )u(x , t ) dxdt. (77)
310 7. Brittle Fracture: The Continuum Approach
If
G(k, ω) = eikx −iωt G(x , t ) dxdt, (78)
Therefore, σ − describes the stresses along the fracture faces, while σ + is an, as
yet unknown, function. u,− on the other hand, is an unknown function along
the fracture faces and vanishes ahead of its tip. Using Eqs. (77)–(79), we write,
G ∗ u = σ , which after Laplace–Fourier transforming yields,
G(k, ω) u(k, ω) = σ (k, ω). (81)
Using Eq. (79) we obtain
G− (k, ω) u(k, ω) = G+ (k, ω) σ (k, ω) (82)
which, after inverting back to real space, yields
G+ ∗ σ = G− ∗ u. (83)
One can show that for x < l(t), G+ ∗ σ + = 0. Suppose that x > l(t). Since σ +
is zero behind the fracture, the integral
G+ ∗ u + = G+ (x − x , t − t )σ + (x , t ) dxdt (84)
is zero for x < l(t ). The only case for the integrand to be non-zero is for x > l(t ),
7.5. The Equation of Motion for a Fracture in an Infinite Plate 311
and similarly
G− (x − x , t − t )u− (x , t ) dxdt = 0, x > l(t). (87)
Therefore
1
G+ (x, t) = √ δ(t − x/c)#(x). (102)
πx
By a largely similar analysis we find that
d #(x)
(G+ )−1 (x, t) = δ(t − x/c) √ . (103)
dx πx
Having calculated (G+ )−1 (x, t), we can now find the stress intensity factor
KI I I (l, t). From Eq. (93) we find that
√ d #(x1 )
K(v) = π ε δ(t1 − x/c) √ #[l(t) + ε − x1 − l(t − t1 )] dx1 dt1
dx1 π x1
√ d #(x1 )
= ε √ #[ε/(1 − v/c) − x1 ] dx1 . (104)
dx1 π x1
Since only very small x1 are important, we find that
!
√ #(x1 )
K(v) = ε √ δ[ε/(1 − v/c) − x1 ] dx1 = 1 − v/c. (105)
π x1
Similarly,
√ #(x1 )
K̃(l, t) = − 2 δ(t1 − x1 /c) √ σ [−l(t) − x1 , t − t1 ] dx1 dt1
π x1
√ #(x1 )
=− 2 √ σ [l(t) − x1 , t − x1 /c]dx1 . (106)
π x1
In particular, when σ − does not depend on the time and, σ (x) = σ0 #(x), one
obtains
√ √
K̃ = −(4/ 2π )σ0 l. (107)
The minus sign arises because the stresses ahead of the fracture tip always act
against those applied on the fracture faces. Note that Eq. (104) reduces to unity
when v → 0, implying that in the case of time-independent loading, K̃ is indeed
the stress intensity factor one would have had if the fracture had been static at l
for all times. For the propagating fracture, we obtain
!
KI I I = 1 − v/c K̃[l(t), σ0 ]. (108)
We now compute the stress singularity that would have developed had we had
a static fracture of length l(t) at time t, and multiply the result by a function of the
instantaneous velocity. We should emphasize that all details of the history of the
crack motion are irrelevant; only the velocity and loading configuration are needed
for determining the stress fields sufficiently close to the tip. As a consequence, one
can use Eq. (75) to determine the energy flow to the tip of the crack:
K̃ 2
H = v(1 − v/c) . (109)
2αµ
314 7. Brittle Fracture: The Continuum Approach
The rate at which energy enters the tip of the fracture must be equal to v(v). There
is nothing to prevent the fracture energy from being a function of the velocity, but
the notion of local equilibrium, which has prevailed until now, strongly suggests
that should be a function of v alone. Therefore
K̃ 2 (l)
(v) = (1 − v/c) , (110)
2αµ
which, after rearranging and using Eq. (106), yields
π µ !
= (1 − v/c)(1 + v/c). (111)
4lσ02
If we define
π µ
l0 = , (112)
4σ02
Eq. (111) is rewritten as
l0 !
= (1 − v/c)(1 + v/c). (113)
l
7.5.2 Mode I
The preceding analysis can also be carried out for thin plates under tension. Al-
though all steps of the analysis proceed as before, it is not possible to derive
simple analytical expressions. This case has been discussed in detail by Freund
(1990) who finds that the energy flux per unit length extension of the fracture, to
an accurate approximation, is given by
(1 − v/cR )K̃ 2 (l)
H(v) = (v) = , (114)
2λ̃
where λ̃ is a Lamé constant defined by Eq. (20). Rearranging Eq. (114) yields
Y (v) v
=1− , (115)
K̃ 2 (l)(1 − vνp2 ) cR
where K̃ is still given by Eq. (106), using σyy on the x-axis for σ . In the case of
time-independent loading described by σ (x) = σ0 #(x), one obtains
l0 v
=1− , (116)
l cR
with
π λ̃
l0 = . (117)
4σ02
Equation (116) is now written in the following form
v = cR (1 − l0 / l), (118)
7.5. The Equation of Motion for a Fracture in an Infinite Plate 315
which is nothing but Eq. (7) obtained by the scaling analysis, with the difference
that and hence l0 can depend strongly upon the crack velocity v. Hence, seem-
ingly large differences between the predictions of the theory and experimental data
are due to nothing more than assuming that l0 is a constant!
What are the practical implications of Eq. (10) for the design of experiments? As
discussed by Fineberg and Marder (1999), one may consider three experimental
situations. (1) One for which the assumptions of the theory hold well. (2) A second
experiment in which the theoretical assumptions are satisfied in an approximate
way, while (3) in the third experiment the assumptions clearly fail. The three cases
are as follows.
(1) A thin plate has a fracture running half-way through, and driven by wedging
action in the middle. For times less than that needed for sound to travel from
the point of loading to the material’s boundaries and back to the tip of the
fracture, all the assumptions of the theory are satisfied.
(2) A thin plate has a long fracture as before, but with uniform static stresses σ∞
applied at the outer boundaries, and the faces of the fracture being stress-free.
This problem is equivalent to one in which the upper and lower outer bound-
aries are stress-free, but uniform stresses −σ∞ are applied along the fracture
faces. The equivalence is due to the fact that an uncracked plate under uni-
form tension σ∞ is a solution of the equations of elasticity, so this trivial static
solution can be subtracted from the first problem to obtain the second equiv-
alent one. However, in the new problem, stresses are applied to the fracture
faces all the way back to the left-hand boundary of the material. Therefore, the
problem must be mapped onto one in which stresses are applied to the faces
of a semi-infinite fracture in an infinite plate, but the correspondence is only
approximate.
(3) Consider now a semi-infinite fracture in an infinitely long strip, shown in Fig-
ure 7.4, which is loaded by displacing each of its boundaries at y = ±w/2 by
a constant amount δ. Far behind the tip (as x → −∞), the fracture relieves
all the stresses within the strip. Far ahead of the tip (as x → +∞), the mate-
rial is unaffected by the fracture with the stress field being linear in y. Thus,
the energy per unit extension far ahead of the fracture has a constant value,
2Y δ 2 /[w(1 − νp2 )], where Y and νp are, respectively, the usual Young’s mod-
ulus and Poisson’s ratio of the material. The translational invariance of the
system along the x-direction implies that, for a given δ, the fracture should
showed that the condition KI I = 0 has the following consequences for fracture
propagation. Consider an initially straight fracture, propagating along the x-axis.
The components σxx and σyy of the stress field are given by
KI
σxx = + σ + O(r 1/2 ), (121)
2π r 1/2
KI
σyy = + O(r 1/2 ). (122)
2π r 1/2
The constant stress σ is parallel to the fracture at its tip. If σ > 0, any small
deviations from straightness cause the fracture to diverge from the x-direction,
whereas if σ < 0 the fracture is stable and continues to propagate along the x-axis.
Yuse and Sano (1993) and Ronsin et al. (1995) conducted experiments described in
Section 6.12 by slowly pulling a glass plate from a hot region to a cold one across
a constant thermal gradient. The velocity of the fracture, driven by the stresses
induced by the non-uniform thermal expansion of the material, follows that of
the glass plate. At a critical pulling velocity, the fracture’s path deviated from
straight-line propagation and developed transverse oscillations. This instability is
completely consistent with the principle of local symmetry: The crack deviates
from a straight path if the stress σ in Eq. (121) is positive. The wavelength of
the ensuing oscillations has also been computed numerically (Adda-Bedia and
Pomeau, 1995).
with those predicted by variants of this criterion, although the same criteria predict
critical velocities for crack branching that are nearly identical to Yoffe’s prediction,
namely, 0.61cR . Adda-Bedia and Ben Amar (1996), for example, proposed that
one should draw contours of constant principal stress and search for points where
these contours are perpendicular to lines drawn from the crack tip, along which
the fracture travels. This criterion predicts the existence of two critical speeds.
The first is the velocity at which the fracture must choose between three possible
directions, whereas the second critical velocity is one at which the fracture must
choose between five possible directions. Although this criterion is plausible, there
is no experimental evidence indicating that this is in fact the preferred criterion.
The branching angles that are predicted by such criteria are not significantly dif-
ferent from those determined by the following condition, which is a type of static
condition. Consider the stress field formed ahead of a single propagating fracture,
from which one can compute the trajectories that satisfy the quasi-static condi-
tion, KI I = 0 (Kalthoff, 1972; Parleton, 1979). The angle that is determined by
this trajectory at a distance rc from the fracture tip, where rc is the typical size
of the cohesive zone, is in good quantitative agreement with the experimental
observations.
Figure 7.5. Comparison of the predictions of linear continuum fracture mechanics (the
curve) with the experimental data (symbols) of Kobayashi et al. (1974).
field within the material. In their experiments strain gauges, with a temporal reso-
lution of about 1 µs, were placed throughout the sample to measure the temporal
behavior of the stress field surrounding a fracture at times immediately following
its arrest. Their data agreed with a prediction of Freund (1990), that the stress field
at a point directly ahead (behind) the fracture should reach its equilibrium value
(to within a few percent) as soon as the shear (Rayleigh) wave front passes.
However, the same type of favorable comparison between the theory and ex-
periment does not exist at high fracture velocities, and in fact experiments often
seem to disagree with Eq. (14). As an example we show in Figure 7.5 the data
of Kobayashi et al. (1974) with PMMA and compare them with the theoretical
predictions. Although the theory predicts that if the fracture energy is not a strong
function of the velocity, the fracture would smoothly accelerate from rest to the
Rayleigh wave speed cR , Kobayashi et al.’s data do not confirm this prediction:
After the fracture initially accelerates rapidly, it becomes increasingly sluggish and
eventually reaches a final velocity well below cR . However, if we suppose that
the fracture energy is a function of the velocity, and specify in Eq. (118) [using
Eqs. (114)–(116)] that l0 is defined in terms of the minimum energy (0) at which
fracture propagation first happens, one obtains instead of Eq. (118):
(v) l0
v = cR 1− . (123)
(0) l
Therefore, if the fracture energy (v) increases rapidly with the velocity v, one can
obtain practically any functional dependence of the velocity on the fracture length.
One can also interpret Eq. (123) as a way of extracting the velocity dependence of
fracture energy from measurements of v. However, validation of the theory cannot
be accomplished without an independent measurement of the fracture energy ,
although even such validation would provide no fundamental explanation of the
origin of any measured velocity dependence of the fracture energy.
7.7. Comparison with the Experimental Data 321
Figure 7.6. Experimental data (triangles) of Sharon and Fineberg (1999) and their
comparison with the theoretical predictions (rectangles).
where c is a wave speed, m is the mass, and m2 δ is the applied force. The fracture
is essentially moving along the center line of a strip of finite width. The traction
applied to the fracture surface was assumed to be given by
. .
∂u .. ∂ 2 u̇ ..
µ . = σc (u) − η 2 . , (126)
∂y y=0 ∂x .
x=0
where µ is an elastic modulus (for example, the shear modulus), and σc is the
cohesive stress acting between the open fracture faces. The cohesive stress was
taken to be σc = σy for 0 ≤ u(x, 0, t) ≤ δc , and σc = 0, otherwise, where σy is the
yield stress, and δc is obviously the range of the cohesive force; note that u(x, 0, t)
is just the fracture-opening displacement. The second term on the right-hand side
of (126) is a viscous damping stress which acts on the fracture surface. The two
spatial derivatives preserve reflection and translational symmetries, and the time
derivative in (126) breaks time-reversal symmetry in order to produce energy
dissipation. The most interesting prediction of the model was a relation between
the velocity of fracture propagation and the externally applied stress, given by
$ %12
v/c K
, (127)
[1 − (v/c)2 ]3/2 Ke
which is valid for 1 !K/KG (w/δc )1/6 (σy /µ)1/6 , where w is the width of
the system, and KG = 2σy δc /µ. Here K is the stress-intensity (more precisely,
the strain-intensity) factor associated with the applied force, and
$ %1/12 $ %2/3
1/3 ηcR σy
Ke = (6δc ) . (128)
µ µ
The surprising aspect of these predictions is the unusual exponent 1/12. If Ke >
KG , then the fracture velocity v jumps from very small values to values near cR
as K passes through Ke , and therefore Ke plays the role of an effective Griffith
threshold at which the fracture makes a sharp transition from slow motion to rapid
propagation. Whether such predictions can be observed in an actual experiment
remains an open question. The dissipated energy is η(∂ u̇/∂x)2 , and assuming that
this energy is converted to heat, then the corresponding temperature rise T will
be
#
K 2 µ3 v
T , (129)
3Cp η
where Cp is the specific heat of the material.
How do the thermal effects in the cohesive zone influence our basic understand-
ing of the fracture process? Since the fracture energy is an input into the theory of
linear continuum fracture mechanics, neither the large temperature rise observed
within the cohesive zone nor its cause(s) have any effect on the predictions of the
theory. This is true so long as the heat dissipation is localized within the cohe-
sive zone and does not spread out throughout the material. Otherwise, the entire
rationale behind Eqs. (91)–(93) would be invalid. Thus, the total fracture energy
7.8. Beyond Linear Continuum Fracture Mechanics 325
surface is a self-affine fractal (see Section 1.3), and studies of aluminum alloys,
steel, ceramics and concrete indicated (Bouchaut et al., 1990, 1991; Måløy et al.,
1992) that the local width w of the fracture surface scales as
w ∼ α , (130)
where is the scale of observation within the fracture plane, and α (which is usually
the same as the Hurst exponent H defined in Chapter 1) is the roughness exponent.
Characterization of rough surfaces and measurement of the associated roughness
exponent α were discussed in Chapters 1 and 6. As discussed in Section 6.14.1,
it appears that for both quasi-static and dynamic fracture a universal roughness
exponent, α 0.8, is obtained for > ξc , where ξc is a material-dependent length
scale (Daguier et al., 1996, 1997). For < ξc a different roughness exponent,
α 0.5, has been measured (Milman, 1994). Narayan and Fisher (1992) inter-
preted α 0.5 as being the result of a crack front pinned by microscopic material
inhomogeneities in very slow fracture.
As already explained in Chapter 6, the apparent length-scale dependence of the
roughness exponent α may also be explained in another way based on the velocity
of fracture propagation.According to Bouchaud and Navéos (1995) (and somewhat
similar to the argument of Narayan and Fisher, 1992), one must distinguish between
quasi-static (slow) and rapid fracture. In the former case, corresponding to small
length scales, one may obtain a roughness exponent close to 0.5, whereas rapid
fracture, which corresponds to large length scales, leads to α 0.8. Bouchaud
and Navéos (1995) thus argued for the existence of a length scale ξqs , such that
for < ξqs one is in the quasi-static fracture regime and thus a low roughness
exponent, while at length scales ξqs rapid fracture is dominant and therefore
one should obtain α 0.8. As shown by Daguier et al. (1997), ξqs depends on the
velocity of fracture propagation, and thus should decrease as the velocity increases.
If this picture of fracture is correct, then models that are based on minimum energy
surfaces are in the quasi-static class. Bouchaud and Navéos (1995) also showed
that the data for both cases can be expressed by the following equation
hmax
= A1 + A2 r α−αms , (131)
r αqs
where hmax is the same as before, αqs is the roughness exponent corresponding
to the quasi-static limit, α is the universal roughness exponent corresponding to
rapid fracture, αms is the roughness exponent of minimum energy surfaces, and
A1 and A2 are two constants.
We should point out that, despite the considerable effort that had gone into
understanding the properties of self-affine fracture surfaces, up until recently, there
was little discussion of the fact that, in many of the experiments in which a non-
trivial roughness exponent had been measured for a fracture surface, the typical
length scales where the scaling behavior had been observed were several orders of
magnitude smaller that the typical sample size. For example, the largest length scale
observed in measurements performed on soda-lime glass (Daguier et al., 1997) was
of the order of 0.1 µm, which is well within the mirror regime. Thus, in the context
7.8. Beyond Linear Continuum Fracture Mechanics 329
where ξ(y) = By 1/z depends on the distance to the initial notch y and corresponds
to the crossover length along the x-axis below which the fracture surface is self-
330 7. Brittle Fracture: The Continuum Approach
affine with a local roughness exponent αloc . The quantity z is the dynamic exponent
for rough surfaces that was already introduced in Section 1.5.
The scaling laws (132) indicate that along the y-axis the roughness develops ac-
cording to two different regimes: For large length scales [ ξ(y)], the roughness
grows as h ∼ y α/z , where α is called the global roughness exponent, whereas
for small length scales [ ξ(y)] the roughness growth is characterized by the
exponent (α − αloc )/z. Unlike the local roughness exponent, the global exponent
α, as well as the dynamic exponent z and the prefactors A and B are material
dependent, and hence non-universal. Thus, despite exhibiting universality in the
transverse direction, roughening in the longitudinal direction is material dependent.
An important consequence of scaling laws (132) is that, when the global sat-
uration occurs, i.e., far from the notch for y ysat [where ysat = (L/B)z ], the
magnitude of the roughness is not only a function of the window size but also
of the system size L, since in this case, h(, y ysat ) Aαloc Lα−αloc . It is
for this reason that scaling laws (132) are viewed as anomalous because in the
conventional scaling of rough surfaces that were described in Section 1.5 one has
αloc , if ξ(y),
h(, y) A (133)
ξ(y) , if ξ(y).
α loc
If fact, the scaling laws (132) and (133) become equivalent only if we take the
global roughness exponent α to be equal to the local exponent αloc .
These anisotropic scaling laws have important implications for the Griffith
criterion which will be described shortly.
the two can be related, since the minimal energy surfaces, such as those obtained in
the random-bond Ising model, seem to have little, if anything, to do with fracture
of a material. Nevertheless, Hansen et al. (1991) suggested, and Räisänen et al.
(1998) confirmed by extensive numerical simulations, that the roughness exponent
of the two types of surfaces in 2D are the same. In particular, Räisänen et al. (1998)
used a scalar approximation to model fracture of a brittle material—the random
fuse model described in Section 5.2—to provide strong numerical evidence for
this equality. However, in 3D the scalar quasi-static fracture model was found to
be rougher than the minimal energy surfaces.
14◦ in Homalite, 15◦ in polycarbonate (Ramulu and Kobayashi, 1985) and about
18◦ in steel (Anthony et al., 1970), all of which were measured for materials that
were under pure uniaxial tension.
Figure 7.8. The dependence of the fracture energy on the fracture velocity v, for (top
row, left to right) AISI 4340 steel (Rosakis et al., 1984), and PMMA (Sharon et al., 1996).
The bottom row shows the rescaled data, where = KI /KI c , and cR is the Rayleigh wave
speed.
Figure 7.8 does in fact reflect the view of Dally (1979) who studied extensively
dynamic fracture in amorphous polymers, and in steels. According to him,
(1) the proper way to characterize a dynamic fracture experiment is through pre-
senting the data by two dimensionless numbers which are v/c, the ratio of the
fracture velocity and a wave speed, and = KI /KI c , the ratio of the dynamic
stress intensity factor and its critical value at the fracture onset. The relation
v/c = f () contains most of the information about the dynamics of fracture.
(2) The energy needed for fracture of brittle amorphous materials increases steeply
past a critical velocity, where the straight-line fracture becomes unstable to
frustrated branching events.
We will come back to these points later in this chapter.
the surface is rough, and thus the stress-field singularity is the usual Griffith’s
singularity, −1/2 . Equating (137) and (141) leads to α = 2 − 2ζ (yielding
the Griffith’s result, ζ = 1/2, when α = 1, i.e., when the fracture surface is
smooth). Thus, rougher fractures, i.e., those with smaller α, lead to a more
singular stress field.
(2) In the second regime, hmax ξ or ξ ξ ∗ . In this case the slope of the surface
over the entire fractal domain is larger than one, resulting in a spiky regime,
and hmax /ξ is a measure of this spikiness. Near the tip of the fracture (r < ξ )
the stress field is characterized by the exponent ζ = 12 (2 − α).
However, the above considerations are valid when the anisotropy in the growth
of rough fracture surfaces is not taken into account. As described in Section 7.8.7,
the height fluctuations in the longitudinal and transverse directions exhibit distinct
scaling properties that are characterized by Eqs. (130) and (132). In particular,
one has an anomalous, size-dependent scaling in the saturation regime, which
must be taken into account if one is to generalize the Griffith criterion for the
onset of fracture. Based on these scaling laws, Morel et al. (2000) proposed a
modified form of the Griffith criterion. To understand their proposal, consider a
semi-infinite linear elastic material of thickness L that contains an initial crack
at position a and in Mode I (i.e., under a uniaxial stable and low tension). In
the zone where the roughness of the fracture surface grows, i.e., for a ysat
[where ysat = (L/B)z defined in Section 7.8.7], the critical energy release rate Hc
during fracture propagation (which, in Griffith’s approach, is set to be equal to the
energy required for generating the corresponding free surfaces at the microscale;
see above and Section 6.7) is given by
2
AB α−αloc
c = 2s 1 + 1−α
a 2(α−αloc )/z , a ysat , (142)
0 loc
where 0 is the lower cut-off for the length scale over which the fracture surface is
a self-affine fractal, i.e., 0 is the characteristic size of the smaller microstructural
element which is relevant for the fracture process, and s is the specific surface
energy that characterizes the resistance of the material to fracturing. The quantities
A and B and the exponents α and αloc were already introduced and discussed in
Section 7.8.7.
On the other hand, when the crack increment is large (i.e., a ysat ), which
corresponds to the saturation state of the roughness, one has
2
A
c = 2s 1 + 1−α
L2(α−αloc ) , a ysat , (143)
0 loc
implying that the energy c is independent of a, but depends on the linear size
L of the sample, an important characteristic of brittle fracture of heterogeneous
materials. Equation (143) indicates that the size effect gives rise to two asymptotic
behaviors which are, c ∼ 2s and c ∼ Lα−αloc . The crossover between the two
338 7. Brittle Fracture: The Continuum Approach
1−α
occurs at a length Lco = (0 loc /A)1/(α−αloc ) . Hence, for L Lco the fracture
surface is shallow, and there is no size effect, c 2s . In this case, the classical
results of linear continuum fracture mechanics are applicable to fracturing of the
material. However, for L Lco one has a power law
c ∼ Lα−αloc > 2s . (144)
Equation (144) was found to agree with the experimental data for wood (Morel et
al., 1998). Note that, if the anomalous scaling is neglected, and the fracture surface
is described by scaling laws (133), then
2
A
c (a) 2s 1 + 1−α
, (145)
0 loc
that is, there is no dependence on the size of the material, which is the case for
purely elastic brittle materials.
2500
Cl
2000
√2Ct
Crack Speed (m/s)
1500
Ct
CR
1000
500
Notch Tip
0
0 20 40 60 80 100 120
Crack Length (mm)
Figure 7.9. Supersonic crack propagation velocity in Homalite-100 (after Rosakis et al.,
1999).
in Section 6.11.2.3 was used for recording the stress field near the propagating
fracture. The sample was subjected to asymmetric impact loading with a projectile
at 25 m/s, and sequences of isochromatic fringe patterns were recorded around a
shear fracture as it propagated along the interface between the two Homalite halves.
Crack tip speeds were measured independently from crack length history. Figure
7.9 shows the speed of the propagating crack versus the crack length. Initially,
the crack tip speed is close to the shear wave speed of Homalite-100, beyond
which it accelerates and becomes intersonic. Thereafter, it continues to accelerate
up to the plane stress dilatational wave speed
√ of the material, then decelerates and
approaches a steady-state value of about 2ct . As mentioned above, the speeds
between cR and ct are in the FVZ.
Observations of fast shear rupture during earthquakes have also provided the
impetus for a considerable amount of theoretical work. We already mentioned in
Section 7.7.1.1 the theoretical work of Langer (1992) which predicted the possi-
bility of supersonic fracture propagation. Theoretical analysis of Andrews (1976)
had already shown that a shear
√ fracture can have a terminal velocity either less than
cR or slightly greater than 2ct , depending on the cohesive strength of the fault
plane ahead of the fracture. Burridge et al. (1979)√ concluded from their theoreti-
cal analysis that the crack speed regime ct < v < 2ct is inherently unstable for
dynamic shear crack growth. Broberg (1989) showed that the crack speed regime
cR < v < ct is forbidden for both opening and shear mode cracks, a result that
was mentioned above. He also showed that the regime ct < v < c√ l is forbidden for
opening mode cracks only. Finally, Freund (1979) showed that 2ct is the only
speed permissible for a stable intersonic shear crack.
340 7. Brittle Fracture: The Continuum Approach
The existence of crack growth with a speed larger than cR has also been con-
firmed by large-scale molecular dynamics simulations of dynamic fracture. These
simulations will be discussed in Chapter 9.
Figure 7.10. Typical measurements of velocity (in m/s) of a fracture tip as a function of
its length in PMMA. The fracture velocity initially jumps to 150 m/s, and then accelerates
smoothly to the critical velocity vc (dotted line), beyond which strong oscillations set in.
The times are in µsec (after Fineberg and Marder, 1999).
342 7. Brittle Fracture: The Continuum Approach
smooth. Instead, one has rapid oscillations in the fracture’s velocity which increase
in amplitude as v does. On the other hand, Hauch and Marder (1999) carried out
experiments in which the energy available per unit length decreased slowly through
the length of the sample. In both PMMA and Homalite-100, fractures decelerated
gradually to zero velocity, supporting strongly the notion that initial trapping, rather
than any intrinsic dynamical effect, is responsible for the velocity jumps, such as
those in Figure 7.10, which are always seen when fractures begin to propagate.
Indeed, in the case of glass, it is possible to prepare very sharp initial cracks so
that their propagation can begin gradually and then continue steadily at velocities
that are only a small fraction of the Rayleigh wave speed cR .
The next question is whether the velocity oscillations are random fluctuations
or are periodic in time. A careful examination of the oscillations indicate that,
although they are not completely periodic, a well-defined time scale does exist with
a value that, in the case of PMMA, is typically between 2 and 3µs. Moreover, in
experiments in which the fracture accelerates continuously, the location of the peak
of the power spectrum of the data in the frequency domain is constant, although
the velocity varies by as much as 60% of its mean value. As Figure 7.10 also
indicates, there is a critical velocity vc beyond which the fracture velocity begins
to oscillate. Many experimental observations indicate that vc is independent of the
sample geometry and thickness, and the applied stress. The value of the critical
velocity for PMMA is about 0.36cR which, when surpassed, results in oscillations
in the fracture velocity and an increase in the fracture surface area.
Figure 7.11. The root mean square values of the surface heights (in µm) as a function of
the mean fracture velocity (in m/s) in PMMA. Different symbols are for various stresses
and sample geometries (after Fineberg et al., 1992).
Figure 7.12. Mean branch length (in µm) as a a function of the mean fracture velocity
in PMMA. The critical velocity is about 340 m/s (the data are from Sharon and Fineberg,
1996, and Sharon et al., 1996).
increases sharply with the mean velocity of the crack. When the velocity reaches a
value of about 1.7vc , the pattern of the growing branches becomes coherent across
the entire thickness of the sample. At still higher velocities, macroscopic branching
occurs. This phenomenon represents a crossover from a 3D behavior to a 2D one.
The second method for quantifying the coherence of the microbranches, which
also helps to further quantify the crossover between the 3D and 2D behavior, is
based on measuring the ratio of the total amount of fracture surface produced by the
crack and its branches located at the sample faces, and that produced at the center
of the sample. Sharon and Fineberg (1996) found that the difference in surface
production between the outer and center planes decreases continuously until the
fracture velocity is about 1.65vc , at which the ratio approaches 1, indicating that
microbranch production across the sample is homogeneous. These results are also
supported by the experiments of Boudet et al. (1996) on PMMA that indicated
that both the sound emissions and surface roughness diverge as the mean fracture
velocity approached 1.7vc , hence suggesting that a second transition may occur at
v ∼ 1.7vc . As the divergence of surface roughness is an indication of macroscopic
branching, the crossover from 3D to 2D may be considered to be a sufficient
condition for macroscopic branching to occur.
in order to explain how the apparent stress singularity actually joins smoothly a
region around the fracture tip where all the fields are finite.
Figure 7.13. Schematics of the cohesive zone model of Barenblatt and Dugdale. The faces
of the fracture are pulled apart by a cohesive stress σc until the faces are separated by a
critical distance lc . The fracture moves from left to right.
7.11. Models of the Cohesive Zone 349
where the negative sign is due to the fact that the cohesive zone is pulling the
fracture faces together and cancelling out the positive stress intensity factor which
is being generated by other forces outside the fracture. Substituting Eq. (148) into
(73) and using Eq. (147) yield
$ %
1 − νp2 2 8L
H= AI (v) σc = lc σc , (149)
Y π
the densities in the gas and liquid phases that are in equilibrium with each other. This
concept is also well-defined for geometrical models, such as the percolation model
for which the order parameter represents the fraction of the uncut bonds or sites,
where the cut bonds represent the “defects.” Hence, at the percolation threshold pc ,
where the geometrical connectivity of the system is lost due to the presence of too
many defects, the order parameter is zero, slightly above pc is very small, while
far from pc the order parameter is nearly unity, since in this region the defects
are too few. In a similar spirit, the order parameter for dynamic fracture should be
related to the concentration of point defects in the material, hence characterizing
local order. In this formulation, the order parameter is (similar to the percolation
model) unity outside of the propagating fracture, but zero inside the crack where
all the atomic bonds have been broken. On the crack surface, the order parameter
varies continuously between 0 and 1, on length scales that are much larger than the
interatomic distances. This would then justify use of a continuum formulation of
dynamic fracture propagation, in which case one would need an equation for the
order parameter that couples it to the equation for the elastic deformation, hence
the name two-field models. The advantage of formulating the problem in terms of
an order parameter and coupling it to the displacement field is that, by allowing
the order parameter to vary in the cohesive zone, the stress singularity at the tip of
the fracture is avoided, hence removing one main deficiency of continuum fracture
mechanics.
One such two-field model was developed by Aranson et al. (2000). They focused
on 2D materials in Mode I fracture, and represented the elastic deformation of an
amorphous material by the usual wave equation, coupled to a term that represents
viscous damping:
$ %
∂ 2u ∂u
ρ0 2 = η∇ 2 + ∇ · σ, (150)
∂t ∂t
where the first term on the right-hand side accounts for viscous damping with η
being the viscosity, and ρ0 is the material’s density which is taken to be unity. The
stress tensor σ is related as usual to the strain tensor , except that their relation
now contains a term involving the order parameter P. This relation, in component
form, is given by
$ %
Y νp ∂P
σij = ij + I δij + a1 δij , (151)
1 + νp 1 − νp ∂t
where a1 is a constant, I is the trace of the strain tensor, and the rest of the notations
are as before. One must take into account the effect of the material’s weakening by
fracture which reduces the Young’s modulus Y . Therefore, Aranson et al. (2000)
assumed that, Y ∼ Y0 P, where Y0 is the initial Young’s modulus. In Eq. (151) the
term that couples the stress and strain tensors to the order parameter accounts for
the hydrostatic pressure that one must apply to the material in order to generate
new defects. Although one might be tempted to interpret this term as being due
to the material’s thermal expansion, such identification would be erroneous since,
as discussed in Chapter 6 and earlier in this chapter, during fracture propagation
7.11. Models of the Cohesive Zone 351
the temperature at the tip of the crack will be high and therefore it is unlikely that
the tip will be in thermal equilibrium. Note that for P = 1—the area outside the
fracture surface—one has the usual equations of elasticity, while for P = 0 the
dynamics is trivial as nothing is happening inside the fracture.
The next step is to develop an expression for the order parameter. Aranson et
al. (2000) assumed that P is governed by purely dissipative dynamics. As such,
the order parameter may be derived from an free-energy functional H,
∂P δH
=− , (152)
∂t δP
which is the standard practice in thermodynamics. In the theory of phase transitions
one has (see, for example, Landau and Lifshitz, 1980)
H= a2 |∇P|2 + Hp (P) dxdy, (153)
Figure 7.14. Fracture velocity v, normalized by the Rayleigh wave speed cR = 926 m/s in
PMMA, versus dimensionless energy H/Hc . Open circles correspond to stable propagation,
crosses to unstable propagation, while diamonds are experimental data of Sharon et al.
(1996). The inset shows the curvature for unstable propagation at ρ = 0.5, with the arrows
indicating the progression of time (after Aranson et al., 2000).
compares the predictions of the model for the crack velocity v in PMMA with
the experimental data of Sharon et al. (1995, 1996). The crack velocity has been
normalized by the Rayleigh speed cR , and is plotted versus the fracture energy H
normalized by its value at v = 0.2c√R . The parameters used in the simulations were
Y0 = 10, νp = 0.36, and η = 13/ Y0 . For PMMA, the Rayleigh wave speed is
cR = 926 m/sec. The model predicts that, depending on the material’s parameter, a
crack instability develops when its speed varies anywhere from 0.32cR to 0.55cR ,
with the instability manifesting itself as pronounced velocity oscillations, sound
emission from the crack tip and, of course, crack branching, as mentioned above.
The agreement between the predictions and the experimental data shown in Figure
7.14 is quite good, indicating the correctness of the model in having most of the
essential features of dynamic fracture. For a somewhat related model see Karma
et al. (2001).
the finite-element (FE) method, have the closest correspondence with experiments
in brittle amorphous materials. In particular, similar to the experiments discussed
above, these FE simulations produced frustrated crack branching, oscillations in
fracture velocities, and limiting crack velocities below the Rayleigh wave speed
cR . Let us describe and discuss these successful efforts.
The basis of Johnson’s work was the physical fact that the size of the cohesive
zone is not predetermined, but is adaptive and changes in accordance with the frac-
ture’s behavior. Since the main purpose of the work was to investigate material
weakening and the role of the cohesive zone, accurate modeling of the contin-
uum region outside the zone was not essential, and therefore Johnson assumed
the continuum to be linearly elastic. In addition, the material modeled was highly
idealized in the sense that, no viscoplastic flow or other rate-dependent properties
were incorporated in the cohesive zone. A planar stress model was used, and an
initial crack of length a0 = 0.6h was inserted in the system, where h is the length
of the plane. The material in the vicinity of the crack tip was assumed to have a
large number of sites where nucleation of defects, all being of the same type and
having the same size, occurs. The fractures were driven by loading their faces with
a number of different loads. Depending on the applied load, the FE simulations
produced maximum fracture velocities of 0.29cR , 0.44cR and 0.55cR . Moreover,
the simulations predicted that, at the lowest velocities, a fracture would accelerate
smoothly. As the external loading was increased, multiple attempts at microbranch-
ing were observed and, similar to the experiments discussed above, the length of
the attempted branches increased with the loading. Moreover, the experimental
observations of Ravi-Chandar and Knauss (1984a) (see Section 7.8.12) that the
stress intensity factor is not a unique function of the fracture velocity v, once v
exceeds a certain limit, were also reproduced by these FE simulations. None of
these results was dependent on the various parameters of the simulations.
More extensive FE simulations were carried out by Xu and Needleman (1994),
although their model of the cohesive zone was different from Johnson’s, and was
also much more elaborate. The continuum was characterized by two constitutive
relations; (1) a volumetric constitutive law that related stress and strain, and (2)
a cohesive surface constitutive relation between the tractions and displacement
jumps across a specified set of cohesive surfaces that were interspersed throughout
the continuum. The first constitutive law was that for an isotropic hyperelastic solid:
∂Hs
σ PK = , (155)
∂
where Hs is the strain energy density which is given by
1
Hs = : C : , (156)
2
where C is the tensor of the elastic moduli. Here, is the Lagrangian strain, and
σ PK is the so-called second Piola–Kirchhoff stress, given by
σ PK = σ · (F−1 )T , (157)
354 7. Brittle Fracture: The Continuum Approach
Eshelby (1969) derived the following equation for anti-plane solution for arbitrary
fracture propagation:
& !
2 K0
u(x, y, t) = Im x − (tr ) + iy , (163)
π Em
where tr = tr (x, y, t) is a retarded time at which a signal arriving at position (x, y)
at time t was launched at the fracture tip, and satisfies the equation, c2 (t − tr )2 =
[x − (tr )]2 + y 2 . The actual analysis is for a finite body prior to the arrival back
at the fracture tip of waves that are reflected from boundaries or from another
fracture tip, in which case the 2D solution very near the tip is given by
& !
2 K
u(x, y, t) = Im x − (t) + iαy + higher order terms, (164)
π αEm
! !
where α = 1 − v 2 (t)/c2 = 1 − (d/dt)2 /c2 , and K is the instantaneous stress
intensity factor given by,
!
K = K0 1 − v(t)/c. (165)
The corresponding energy release rate E is then given by
!
H = H0 [1 − v(t)/c]/[1 + v(t)/c], (166)
where H0 = K02 /(2M).
Rice et al. (1994) derived the 3D solution as a linearized perturbation about the
2D solutions for a fracture propagating at a steady speed v0 [hence, (t) = v0 t].
Thus, if we use polar coordinates such that, r exp(iθ ) = x − v0 t + iα0 y, the 2D
solution becomes
& ! &2 K √ $ %
2 K0 0 1
u0 (x, u, t) = Im x − v0 t + iα0 y = r sin θ ,
π α0 E m π α 0 Em 2
(167)
which is consistent with that of actual elastodynamics for anti-plane strain, if we
identify Em and c with the shear modulus and shear wave speed. To develop the 3D
solution, one sets x = (z, t) = v0 t + f (z, t), a first-order expansion in about
the 2D results corresponding to a straight fracture ( = 0) propagating along the
x axis with a constant velocity v0 . Thus, the shape of the fracture front can deviate
from being straight. The 3D solution is then of the form,
u(x, y, z, t; ) = u0 (x, y, t) + φ(x, y, z, t) + O( 2 ), (168)
where φ(z, y, z, t) = (∂u/∂)=0 . The singular part of the 3D solution must be of
the 2D character, but now relative to the local direction of fracture propagation,
so that for any we must have
& !
2 K(z, t; )
u(x, y, z, t; ) = Im x − (v0 t + f ) cos γ + iα(z, t; )y + · · ·
π Em α(z, t; )
(169)
7.11. Models of the Cohesive Zone 357
!
where α(z, t; ) = 1 − v 2 (z, t; )/c2 , v(z, t; ) = (v0 + ∂f/∂t) cos γ (z, t; ),
and cos γ (z, t; ) = [1 + (∂f/∂z)2 ]−1/2 , with γ being the angle between the local
normal to the fracture front and the x- axis. It is then easy to show that as r → 0
[i.e., as x → (z, t) and y → 0] one has
& $ %
√ 1 K0 1
lim [φ(x, y, z, t) r] = f (z, t) sin θ , (170)
r→0 2π α0 Em 2
so that φ(x, y, z, t) satisfies the same equation as (161), subject to the stress-
free boundary condition, ∂φ/∂y = 0 at y = 0 if x < v0 t. We also have, by
symmetry, φ = 0 at y = 0 when x > v0 t. In the harmonic case, f (z, t) =
F (k, ω) exp(−ikz + iωt), the solution for φ is written as
φ(x, y, z, t; k, ω) =
&
1 K0
F (k, ω) exp[i(ωt − kz)] exp[−iωv0 (x − v0 t)/α02 c2 ]ψ(x − v0 t, y; k, ω).
2π α0 Em
(171)
Since φ must satisfy Eq. (161), we find that ψ must satisfy the following equation
$ 2 %
∂2 1 ∂2 ∂ 1 ∂ 1 ∂2
+ 2 2 ψ= + + 2 2 ψ = Q2 ψ, (172)
∂x 2 α0 ∂y ∂r 2 r ∂r r ∂θ
where
1/2
|k| ω2
Q(k, ω) = 1− 2 , ω2 < α02 k 2 c2 , (173)
α0 α0 k 2 c 2
1/2
iω α02 k 2 c2
Q(k, ω) = 2 1 − , ω2 > α02 k 2 c2 . (174)
α0 c ω2
Equation (173) corresponds to letting k approach the positive real axis through
Im(k) > 0 and the negative real axis through Im(k) < 0; these approaches are then
taken to be branch-cut portions of the Re(k)-axis where |k| > |ω|/(α0 c). Equation
(174) holds for any direction of approach. Note that the combination α0 c, which
often appears in solutions of fracture propagation problems, has a clear physical
interpretation: It is the speed at which information is transmitted transversely along
the propagating fracture front. That is, two points of the fracture front a distance
z apart do not influence each other before the time delay z/(α0 c).
The solution ψ must satisfy the asymptotic requirement (170) as r → 0.
Any more general fracture perturbation f (z, t) can be represented as a Fourier
superposition, so that
+∞ +∞
F (k, ω) = f (z, t) exp[−i(ωt − kz)] dzdt. (175)
−∞ −∞
358 7. Brittle Fracture: The Continuum Approach
The general solution for φ(x, y, z, t) for any f (z, t) is then given by
& +∞ +∞
r K0 sin( 12 θ ) ∂f (z , t ) c(t − t ) − v0 (x − v0 t)/α02 c
φ(x, y, z, t) =
2π 3 α0 Em −∞ −∞ ∂t (x − v0 t)2 /(α02 ) + y 2 + (z − z )2
!
#[c(t − t ) −
(x − v0 t )2 + y 2 + (z − z )2 ]
× ! dt dz ,
c (t − t ) − (z − z )2 − y 2 − (x − v0 t )2
2 2
(176)
where #[ ] is the Heaviside unit-step function. Once ψ is obtained, φ, and hence
the displacement field u, are also obtained.
One can now derive an expression for the stress intensity factor (and hence the
energy release rate). To do this, it is convenient to replace f (z, t) by (z, t) − v0 t
and ∂f (z, t)/∂t by v(z, t) − v0 in all the expressions. To obtain the first order
perturbation to the stress intensity factor at some location ζ along the z-axis, the
crack front (z, t) is written as
(z, t) = v0 t + [(ζ, t) − v0 t] + {(z, t) − (ζ, t)}, (177)
where the [ ] term describes a 2D perturbation, which is solvable exactly to all
orders by Eqs. (163), (165) and (166), while the { } term corresponds to a 3D
perturbation that vanishes at z = ζ for all t. The stress intensity factor at z = ζ ,
due to small deviations from straightness in other fractures,
√ is determined by
applying to solution (176) for φ the operator limr→0 Em 2π r∂/∂y. The result is
given by
! ! ! 7 ! 8
K(z, t) = K0 1 − v0 /c + K0 1 − v(z, t)/c − K0 1 − v0 /c + K0 1 − v0 /c I (z, t) ,
(178)
with
+∞ t−|z−z |/(α0 c)
1 c(t − t )[v(z , t ) − v(z, t )]
I (z, t) = PV ! dt dz ,
2π −∞ −∞ (z − z )2 [α0 c(t − t )]2 − (z − z )2
(179)
with PV denoting the principal value integral, and v(z, t) = ∂(z, t)/∂t being the
local velocity of the propagating fracture. Therefore, the dependence of the stress
intensity factor on the shape of the fracture front and its deviations from being
straight are expressed in terms of I (z, t). The [ ] term of Eq. (178) is actually
exact for arbitrarily large perturbations of v(z, t), but the { } term is exact only to
first order in the deviation v(z, t) − v0 . The choice of v0 is arbitrary so long as it
is in the range of “first-order difference” from v(z, t).
If we examine the expression for I (z, t), we see that when a segment of the frac-
ture front suddenly slows down relative to neighboring locations along the front, a
reduction in K radiates outward from that segment at speed α0 c. Similarly, when
a segment speeds up, an increase of K is radiated. Such elementary slow-down
and speed-up are due to the encounters of the fracture front with regions of higher
or lower resistance to fracture. Rice et al. also (1994) found that when a straight
fracture front approaches a slightly heterogeneous strip which lies parallel to the
fracture tip along an otherwise homogeneous fracture plane, it may be pinned by
asperities after some advancement into the heterogeneous region, if it is propagat-
7.11. Models of the Cohesive Zone 359
ing with a relatively small velocity. If, however, the velocity is relatively high, the
asperities give way, the fracture front becomes curvy and propagates further into
the bordering homogeneous region, where it recovers a straight-line configuration
through slowly-damped space-time oscillations which, if they are in response to
spatially-periodic heterogeneities, decay as t 1/2 with time. Such a slow decay sug-
gests that the configuration of a straight fracture front may be sensitive to even
small but sustained heterogeneity in the fracture resistance (i.e., in the material).
Using the results of Rice et al. (1994), Perrin and Rice (1994) showed that a frac-
ture propagating through a heterogeneous material, in which the heterogeneities
are represented as randomly-distributed asperities with which the fracture front
interacts continually, will never reach a statistically steady state. Instead, het-
erogeneities in the fracture energy lead to a logarithmic divergence of the root
mean-squares deviations of an initially straight fracture front. In particular, the
variance V of the deviation of propagation velocity from the mean, was found to be
V ∼ log(2α0 v0 t). (180)
More interestingly, if the material is uniform over the remaining part of the fracture
plane, after the encounter with the heterogeneous portion of the material, the prop-
agating fracture becomes asymptotically (i.e., in the limit t → ∞) straight again.
These predictions suggest that perhaps the roughness of a fracture surface may be
the direct result of a continuous roughening of the surface that is driven by small
heterogeneities within the material. More recently, Willis and Movchan (1995)
and Movchan and Willis (1995) computed the coupling of the energy release rate
to random perturbations to the fracture front in the case of planar perturbations
to the crack in Mode I fracture, and in shear loading. Willis and Movchan (1997)
extended the analysis to the perturbations to the stress intensity factors induced by
a small 3D dynamic perturbation of a propagating, nominally planar, fracture.
Ramanathan and Fisher (1997, 1998) calculated the dynamics of planar pertur-
bations to a tensile crack front and found that, in contrast to the case of the scalar
model for which Perrin and Rice (1994) had obtained logarithmic instability of
the crack front, in Mode I fracture weak heterogeneity of the material can lead to
a non-decaying unstable mode that propagates along the fracture front. They pre-
dicted that this propagating mode occurs in materials having ∂/∂v ≤ 0, where a
constant value of is a marginal case. For ∂/∂v > 0, the propagating mode was
predicted to decay, with the propagation velocity of the new mode being between
0.94cR and cR . These predictions are supported by the numerical simulations of
Mode I fracture in a 3D material with a constant , carried out by Morrissey and
Rice (1998), indicating that the propagating mode is highly localized in space, and
indeed propagates at the predicted velocities.
Ramanathan and Fisher (1997, 1998) and Morrissey and Rice (1998) both
showed that these localized modes lead to linear growth of the root mean-square
deviations of an initially straight fracture with its distance of propagation. They
suggested that this may provide a new mechanism for the roughness produced
by a propagating fracture in materials in which the fracture energy does not in-
crease rapidly with the velocity of a crack. Both the calculations and simulations
360 7. Brittle Fracture: The Continuum Approach
as the fracture begins to propagate, the greatest tensile forces are perpendicular to
its tip and not parallel to it. Therefore, it is difficult to imagine how a fracture can
ever propagate in a straight line.
That Langer and collaborators found their dynamic models of the cohesive zone
to be unsatisfactory may imply that, such models must be replaced by those in
which plastic yielding is distributed across an area, and not restricted to a line.
The two-field continuum models of the type described in Section 7.11.2 represent
progress in this right direction. Another possibility is that calculations of Langer
and co-workers indicate a fundamental failure of the continuum formulation of
the type that they employ, and that the resolution must be sought either at the
atomic or molecular scale (see Chapters 9 and 10), or one should resort to two-
field continuum models that take into account the variations of the order parameter
in the fracture zone.
and hence trigger extensive plastic deformation before the fracture can propagate
by cleavage. However, Argon (1987) showed that the Rice–Thomson-type models,
in which the activation configuration consists of a fully-developed dislocation line,
greatly over-estimate the energy barriers to nucleation of dislocations. This remains
true even if one considers a modified Rice–Thomson-type model developed by
Cheung et al. (1991) in which fracture tip nonlinearity and tension softening were
incorporated.
On the other hand, consider the response of silicon, and many other similar co-
valent compounds and materials, that have very sluggish dislocation mobility, and
hence are in contrast with high-mobility hypothesis and the nucleation-controlled
response of some materials. In such materials, the transition from brittleness to
toughness is governed by the mobility of groups of dislocations that are away from
the tip of the fracture (see, for example, St. John, 1975; Hirsch et al., 1989; George
and Michot, 1993). It is now well-established for both classes of materials that, the
emission of the dislocations from the tip of a fracture occurs preferentially from
specific sites on the tip, and that, in order to guarantee ductile behavior, the entire
fracture front must be shielded from local break-out of the cleavage fracture from
unprotected parts of the fracture front. Thus, it is now widely believed that the
fundamental BTD transition is governed by the behavior of a cleavage crack.
In addition to the experimental studies mentioned above, theoretical analyses
of fracture behavior of Si, carried out by Rice and Beltz (1994) and Xu et al.
(1995), indicate that the activation configuration of dislocation embryo is a double
kink of dislocation core matter. Thus, one may identify two distinct types of BTD
transitions:
(1) In the BCC transition metals, where barrier to kink mobility along the dislo-
cation are low, the BTD transition is governed by the formation of dislocation
embryos at the fracture tip, which then results in a nucleation-controlled
transition.
(2) By contrast, experimental work (see, for example, Yonenaga and Sumino,
1989) and theoretical modeling (Bulatov et al., 1995) suggest that, in semi-
conductors and compounds the kink mobility is hindered by substantial energy
barrier, hence rendering the BTD transition controlled by dislocation mobility
away from the tip of the fracture.
A complete understanding of the BTD transition can be obtained based on atom-
istic modeling of the formation and outward propagation of the dislocation embryo
at the tip of the fracture. Such atomistic modelings are based on ‘molecular dy-
namics simulation that will be described in Chapters 9 and 10. However, atomistic
models provide quantitative predictions for this phenomenon only if accurate po-
tentials for describing the interatomic interactions are available. Several promising
interatomic potentials have been developed over the past decade or so that will
be described in Chapter 9. Alternatively, one may utilize a multiscale modeling
approach—one that combines continuum modeling for the region away from the
fracture tip with atomistic simulations in the tip region—in order to study this phe-
nomenon. This represents a realistic and powerful approach that is rapidly gaining
7.12. Brittle-to-Ductile Transition 363
popularity; Chapter 10 will describe this method. So far as the BTD transition is
concerned, Xu et al. (1995) have already developed a multiscale model for study-
ing this phenomenon. They showed that the energetics of the dislocation embryo
formation on inclined slip planes that contain the fracture tip, when compared with
an additional surface production resistance, is quite unfavorable and cannot ex-
plain the known BTD transition temperatures. Xu et al. conjectured that nucleation
may be more favorable on oblique slip planes, or may occur heterogeneously at
the edges of the fracture front. However, we must realize that, although disloca-
tion nucleation on oblique planes has often been suggested as a likely scenario,
approximate analyses that were based on the Rice–Thomson criterion have led to
estimates of TBTD that are several orders of magnitude larger than the experimental
values.
We note that, although experiments have established the ability of disloca-
tion nucleation at the fracture tip for accounting for the exceedingly sharp BTD
transitions in Si and similar materials, Khanta et al. (1994) questioned this well-
understood fact, and instead advocated an approach based on an analogy with
thermal phase transitions. Specifically, they considered, unlike the more tradi-
tional methods described above, the thermally-induced instability of many small
loops in the presence of an applied stress, and proposed that the creation of many
atomic-size loops by thermal activation induces a temperature-dependent coop-
erative screening effect that enhances the subsequent growth of the loops. This
cooperative effect is completely different from the dislocation shielding of frac-
ture tip stress described above. To develop their theory, they extended the concept
of dislocation screening, originally developed by Kosterlitz and Thouless (1973)
in an entirely different context, namely, 2D phase transitions. In the Kosterlitz–
Thouless (KT) theory, the generation of dislocations (which is an unstable process)
is driven by only thermal fluctuations, without the aid of an applied stress. The
KT transition occurs at a temperature close to the melting temperature, which then
gives rise to a dislocation-mediated melting transition (Nelson and Halperin, 1979;
Young, 1979). In the model developed by Khanta et al. (1994), both the external
stress and thermal fluctuations assist the growth of dislocation loops. The model
then predicts the existence of a KT-type instability, but not a phase transition in
the thermodynamic sense, at a temperature well below the melting temperature,
at a stress level that corresponds to the Griffith threshold that is needed for brittle
fracture propagation. This temperature is then identified with TBTD . If the transi-
tion temperature is zero and the applied load is equal to the Griffith threshold, the
model reduces to the Rice–Thomson model described above. Thus, one advantage
of this theory is that it is applicable to systems that are at a finite temperature, in
contrast with the Rice–Thomson model that is strictly valid for zero temperature.
Despite this success, there is not yet convincing evidence for the role of thermal
fluctuations advocated by Khanta et al. (1994). Indeed, the meticulous experi-
ments of George and Michot (1993), who used X-ray direct imaging of the stages
of evolution of the fracture-tip plastic response, starting from nucleation of crack
tip heterogeneities and followed by very rapid spread and multiplication of dislo-
cation length from such sources, demonstrate clearly the vast numbers of degrees
364 7. Brittle Fracture: The Continuum Approach
Summary
As stated at the beginning of this chapter, it was believed for a long time that there is
a conceptual problem with the continuum mechanical formulation of brittle fracture
of amorphous materials, as its prediction for the terminal velocity of propagating
fractures, i.e., the Rayleigh wave speed cR , had seemed to be experimentally
unattainable (apart from highly anisotropic materials). However, the discussions
of this chapter should have made it clear that the problem persisted not because of
a fault in the continuum mechanics, but because it had not been properly posed.
The correct question should have been about the nature of energy dissipation near
the fracture tip. However, such a problem was not studied for several decades,
because it had seemed natural to assume that, in a sufficiently brittle material,
energy will be consumed mainly for breaking the atomic bonds and generating
new fracture surface, a process that should depend only weakly on the fracture
velocity. However, by loading fractures in differing fashions, greatly-fluctuating
quantities of energy can be forced into the fracture tip. The tip must then find
some mechanism for dealing with the energy not needed to break a minimum set
of atomic bonds. A small fraction of the remaining energy is consumed by such
minor events as phonon emission, after which the tip begins consuming energy by
a sequence of dynamical instabilities, giving rise to ramified networks of fractures
(or broken atomic bonds) on small length scales.
Thus, there is actually no discrepancy between the conventional continuum
fracture mechanics and the experimental observations and data. In a large enough
amorphous material, the fracture-tip instabilities occur within the cohesive zone
where linear continuum fracture mechanics is not even an appropriate theoretical
framework for analyzing the instabilities, let alone predicting them. The finite-
element simulations, models of fracture propagation in 3D, the two-field continuum
models, the lattice models that will be described in Chapter 8, and many precise and
beautiful experiments carried out over the past decade, have now provided us with
a much better understanding of the structure and dynamics of energy dissipation in
the vicinity of the tip of a propagating fracture in a brittle material. It is now clear
that fracture in brittle materials is governed by a dynamic instability that gives
rise to repeated attempts for branching off of the main propagating fracture, hence
preventing the terminal fracture velocity from reaching the Rayleigh wave speed.
8
Brittle Fracture: The Discrete Approach
8.0 Introduction
As discussed in Chapters 6 and 7, theoretical and computer simulation studies of
fracture of materials are usually based on one of the following three approaches.
(1) The first approach formulates the problem using linear continuum fracture
mechanics. This approach, which was described in detail in Chapter 7, allows
one, in many cases, to derive the analytical solution of the problem of frac-
ture propagation in a given material, subject to certain initial and boundary
conditions. If, however, such analytical solutions cannot be derived, then the
governing equations must be discretized by, for example, a finite-difference
or finite-element method and solved by numerical simulations, in which case
the model reduces to a type of discrete or lattice model.
(2) The second approach is based on molecular dynamics (MD) simulation of
fracture propagation which studies the phenomenon at atomic length scales.
Molecular dynamics is a discrete approach in that, the system under study is
represented by a discrete set of atoms connected to one another by atomic
bonds. This approach will be described in Chapter 9.
(3) The third approach is based on lattice models which can be used for both quasi-
static and dynamic fracture phenomena. However, we must point out that there
is a major difference between lattice models of fracture that we describe and
discuss in this chapter and the MD approach to fracture. The difference is due
to the fact that, in MD simulation of fracture breaking of an atomic bond is a
natural outcome of the simulations, whereas in the lattice models described in
this chapter, how or when a bond breaks is an input of the models that must
be specified at the outset. There are, in general, two types of lattice models.
(i) One class of such models is intended for quasi-static fracture. Such mod-
els consist of a lattice of springs or beams, together with a criterion for
nucleation of local microcracks. In these models, each node of the lattice
is connected to only a finite number of other sites (which are usually the
nearest-neighbor sites), and a force balance is written down for each node,
resulting in a set of simultaneous equations that govern the nodal displace-
ments. Unlike the MD method, the nodes of the lattice do not represent the
material’s atoms, nor do the bonds represent the atomic bonds. Instead,
366 8. Brittle Fracture: The Discrete Approach
the lattice models represent a material at length scales much larger than
the distance between two neighboring atoms in the material, and therefore
one does not have to be concerned about developing accurate interatomic
potentials between the atoms, a subject that will be discussed in detail in
Chapter 9.
(ii) The second class of such models are intended for dynamic fracture. This
class of models is itself divided into two subclasses. (a) In one group are
models that represent generalization of the lattice models of quasi-static
fracture. The nodes of the lattice do not represent atoms. Some of such
models contain quenched (fixed in space) disorder, while others have been
developed for fracture of materials with annealed disorder (i.e., one that
may change with the time). (b) The lattice sites in the second group do
represent atoms. However, instead of assuming interatomic potentials be-
tween the atoms, as in MD simulations, one adopts, in a manner similar
to lattice models of quasi-static fracture, a simple force law between the
atoms, one in which the forces rise linearly up to a critical separation be-
tween the atoms, beyond which they abruptly vanish. If the lattice contains
no disorder, then exact calculations can be carried out (see below).
(2) Over the past fifteen years there has been considerable theoretical progress
towards understanding the dynamics of elastic manifolds moving through dis-
ordered media, such as charge density waves (see, for example, Narayan and
Fisher, 1992), fluid-surface contact lines (see, for example, Ertas and Kardar,
1992), and interfaces between two phases, such as those that are encoun-
tered in multiphase flow in a disordered porous medium (see, for example,
Sahimi, 1993b, 1995b), all of which exhibit a sort of non-equilibrium critical
phenomenon close to the onset of motion. Fracture of materials does have
similarities with these phenomena (although it has important differences too)
which have provided the impetus for developing some of the models that were
described in Chapter 7, and those that will be described in the present chapter.
In particular, one is interested to understand the extent of the similarities be-
tween these seemingly different phenomena, so that the possibility of a unified
approach to most, if not all, of them can be explored. Moreover, if such sim-
ilarities do exist, then the knowledge that already exists about some of such
phenomena can be immediately “transferred” into new insight about fracture
phenomena.
To make this point clearer, let us go back to Chapter 7 and recall the essentials
of brittle fracture phenomenon. Suppose that there exists a crack front in a material
and that an external load σ is applied to it. If σ is small, there is no steady-state
motion and the crack front is pinned by the heterogeneities of the material in one
of the many locally-stable configurations. As the external load increases, there are
a series of local instabilities that become larger as σ increases further. At a critical
load (stress) σc the crack front depins and begins to move. In a large enough
system, the transition from the stationary to the moving state exhibits features
of a non-equilibrium dynamic critical phenomenon which, to some extent, are
similar to those of second-order phase transitions, such as the percolation transition
emphasized in this book. For example, the mean velocity v of the moving fracture
just above σc obeys the following power law (Ramanathan and Fisher, 1997):
v ∼ (σ − σc )ζ , (1)
where ζ is a critical exponent which is, hopefully, independent of many micro-
scopic properties of the material. Moreover, in the quasi-static case, as σ increases,
segments of the crack front overcome the local toughness caused by the hetero-
geneities and move forward, causing other segments to jump, thereby triggering
an avalanche which will eventually be stopped by tougher regions. It has been
found that, up to a characteristic length ξ − , the avalanches exhibit a power-law
size distribution, where by size we mean roughly the extent l along the crack front
of an avalanche. This size distribution is given by
P (size > l) ∼ l −κ f (l/ξ − ), (2)
where κ is a characteristic critical exponent. The cutoff length scale ξ − itself obeys
the following power law near σc :
−
ξ − ∼ (σc − σ )−ν , (3)
368 8. Brittle Fracture: The Discrete Approach
where ν − is the critical exponent associated with ξ − . Note that the cutoff length
scale ξ − plays a role similar to ξp , the correlation length of percolation which, as
has been emphasized throughout this book, plays a fundamental role in determin-
ing the length scale over which materials with percolation heterogeneity can be
considered as homogeneous. Moreover, we expect that
σc
l −κ f (l/ξ − ) dσc ∼ l −1 . (4)
0
Just above σc , the fluctuations in the crack velocity are correlated up to a length
scale ξ + which follows another power law given by
+
ξ + ∼ (σ − σc )−ν . (5)
In general, we expect ν − = ν + = ν (see Chapter 3 for examples for which this is
not true). As discussed in Chapters 6 and 7, at the threshold σc the fracture surface
has a self-affine structure with a roughness exponent α, so that the correlation
function C(r) scales as,
C(r) ∼ r 2α . (6)
Finally, the time scale tl that an avalanche of size l lasts is characterized by a
dynamic exponent z, similar to what was defined in Chapter 2:
tl ∼ l z . (7)
Not only are these exponents well-defined, but also satisfy certain scaling relations.
In fact, Ramanathan and Fisher (1997) showed that
ζ = (z − α)ν, ν = (1 − α)−1 , (8)
so that, similar to percolation and other second-order phase transitions, there are
only two independent exponents that characterize this transition. Two-dimensional
(2D) numerical simulations of Ramanathan and Fisher (1997) yielded, z 0.74,
α 0.34, ν 1.52, and ζ 0.34. The estimated α is smaller than the typical
value of the roughness exponent, α 0.8, that, as discussed in Section 7.8.7,
has been reported for several classes of materials. However, MD simulations of
fracture by Nakano et al. (1995), to be described in Chapter 9, indicate that, in
agreement with our discussion in Chapter 7, there may be two regimes of fracture
propagation, characterized by different roughness exponents. Nakano et al. found
that at the initial stages of fracture propagation, when the crack tip moves slowly,
α 0.44, which is reasonably close to the estimate of Ramanathan and Fisher
(1997), while at latter stages when fracture propagation proceeds at relatively high
speeds, α 0.8.
In addition, the lattice models that are described in this chapter have enabled
us to resolve the conflicts between the predictions of linear continuum fracture
mechanics and the experimental observations. In particular, the phenomena of
fracture instabilities, microbranching, and the inability of a propagating fracture
for reaching the Rayleigh wave speed cR (the experimental aspects of which were
8.1. Quasi-static Fracture of Fibrous Materials 369
which is also shown in Figure 8.1. The survival probability ps (the probability
that the bundle does not fail macroscopically) is then, (survival probability of a
1D bundle)L .
The model is physically viable only if the applied stress or strain is shared by
the bonds in a meaningful manner, and thus the issue of load sharing is critical. As
discussed by Duxbury and Leath (1994a), there are two classes of such load-sharing
models which we now describe and analyze.
the democratic model may appear to be difficult but, as pointed out by Sornette
(1989), it can in fact be solved by using the theory of extreme order statistics which
was also used in our discussion of models of electrical and dielectric breakdown
of materials in Chapter 5. The key idea is that, the bundle will not break under an
external load σ if there are k links in it, each of which can withstand a load σ/k. In
other words, if X1;n ≤ X2;n ≤ · · · ≤ Xn;n is the way in which the strengths of the
individual links are ordered, then, if the first k − 1 weakest links fail, the bundle
will resist macroscopic failure under a stress σn ≤ (n − k + 1)Xk;n , because of the
remaining (n − k + 1) links of breaking strength ≥ Xk;n . Therefore, the strength
σn of the bundle is given by
We now search for the strongest subgroup of the bonds. The variables Xk;n are
strongly dependent since they are correlated. However, regardless of the specific
form of F (x), there is a very general result for σn due to Galambos (1978) which
is as follows.
Equation (11) states that the density distribution of the global failure threshold
is Gaussian around the maximum σ = ny with a variance that scales as n, hence
implying that the typical strength of the system increases as σn ∼ n, if n is large.
Although by a naive argument one may predict that σn = nx, where x is the
mean one-link threshold, Eq. (11) shows that σn = ny, with y being in fact signif-
icantly smaller than x, and therefore the naive argument greatly overestimates
the global failure threshold.
The mechanical characteristics of the system under a given applied stress σ < σn
depend upon the history of the system, i.e., on the number and the way the links
have failed as the stress was increased from zero to σ . With each value of σ < σn
we associate an integer m(σ ) with 1 ≤ m(σ ) ≤ n such that
Note that m(σ ) − 2 is the number of links which have failed under a stress ≤
σ/[n − m(σ ) + 2]. Moreover, by definition of F (x),
(m − 2)/n ≤ F [σ/(n − m + 2)] ≤ (m − 1)/n, (14)
which follows from the fact that, for large n, counting the number of links with fail-
ure threshold less than σ/[n − m(σ ) + 2] amounts to computing the cumulative
failure distribution F (x) at x = σ/[n − m(σ ) + 2]. Relations (13) and (14) indi-
cate, roughly speaking, that, as n → ∞, σ/n is increasingly better approximated
by x[1 − F (x)] with
σ
= x(σ ){1 − F [x(σ )]}. (15)
n
Note that Eq. (15), in the limit n → ∞, is a continuous function. It is then not
difficult to show that, for large n, the number of links which have failed under σ
is given by
k(σ ) = nF [x(σ )]. (16)
For large but finite n, σ (x) or x(σ ) is a staircase with plateaux of width decreasing
to zero as n → ∞. The width of each plateau, for a given σ , can be obtained
from (13), since the interval in σ is such that (13) holds with the same integer
m(σ ) = m.
Just before complete failure of the bundle, the total number of failed links is
given by
kn = k(σn ) = nF (x0 ), (17)
implying that a finite fraction of the links fail before global rupture occurs. If
we consider, for example, the (cumulative) Weibull distribution (WD) (see also
Chapter 5),
F (x) = 1 − exp[−(x/λ)m ], (18)
where λ and m are the parameters of the distribution, then
kn
= 1 − exp(−1/m), (19)
n
which for m = 2 yields kn /n = 0.393. For σ ≤ σn , x(σ ) is in neighborhood of x0
and may be expressed as
x(σ ) = x0 − A(y − σ/n)1/2 , (20)
where A is a constant with a value that depends on the shape of F (x). For example,
for the WD, A = [x0 exp(1/m)/m]1/2 . Then, the number of links that have failed
under the stress σ is given by
k(σ )
= F (x0 ) − B(y − σ/n)1/2 , (21)
n
where B is another constant. For example, for the WD, B is given by B =
(mx0 )−1/2 exp(−1/2m). Equation (21) indicates that k increases rapidly as σ →
σn , approaching nF (x0 ) with a square-root singularity.
8.1. Quasi-static Fracture of Fibrous Materials 373
We can thus predict the strain-stress characteristics of the bundle of the fibers.
Suppose that each individual link is made of a brittle material, so that its strain-
stress relation is given by, l = κσl up to its failure point, where l is the strain.
Then,
(1) for σ ≤ σ1 , where σ1 is the strength of the weakest link (the first to fail), all
links are intact and the system has a linear stress-strain characteristic with
slope κ −1 . Note that for the WD, σ1 ∼ λn−1/m .
(2) For σ1 ≤ σ ≤ σn , some of the links have failed, and the system is elastic but
nonlinear, which can be established by the following argument. We see from
Eq. (16) that n{1 − F [x(σ )]} links support the total external stress σ , which
means that the stress per remaining link is given by
σ
σr = = x(σ ). (22)
n{1 − F [x(σ )]}
Thus, for every σr there is a corresponding strain per link r , which is equal
to the strain of the entire bundle of links associated in parallel, and is given by
r = κx(σ ), (23)
done, because failure depends on the largest vacant cluster the statistics of which
are difficult to analyze.
An elegant solution of this problem was developed by Duxbury and Leath
(1994a) (for the solution of the problem in which the stress carried previously
by a failed fiber is shared by its nearest and next-nearest neighbors, see Phoenix
and Beyerlein, 2000). We present a brief description of their solution. With the
cluster-end-load-sharing rule, the bond which suffers the largest stress enhance-
ment is one at the end of the largest cluster of the absent bonds. Under this scenario
then, the survival probability is related to the probability PL (n) that there is no
cluster of vacant bonds of size greater than some prescribed value n. An important
load sharing rule is that, σt = σ (1 + 12 n), where σt is the stress at the tip of the
failed bond. Duxbury and Leath (1994a) calculated PL (n) following a method
proposed by Harlow (1991) in which one identifies the possible endings of a fiber
bundle of length L + 1, and the way by which these endings may be generated from
a bundle of length L. In essence, this method is similar to the transfer-matrix tech-
nique described in Section 5.14.2 of Volume I. Suppose that {1} stands for a present
(unbroken) bond and {0} for an absent (failed) one. If the size of the vacant sites
is restricted to be n, then the bundle endings that are allowed are (1), (10), (100),
(1000· · ·), where the number of zeros in the last probability is n. One now constructs
a transition probability matrix for going from each of these possible configurations
at the end of a bundle of length L to the same endings in a bundle of length L + 1,
by considering the probability of their occurrence. For example, the probability of
going from ending (1) to ending (10) is q, since the probability that the next bond
added is vacant is just q. We define PTL = [p(1) , p(10) , p(100) , · · · , p(100···0) ] as the
probability vector of having the set of possible endings on a fiber bundle of length
L. Then PL+1 is obtained from MPL = PL+1 = ML P1 , where
⎡ ⎤
p p p ··· p
⎢ q 0 0 ··· 0 ⎥
⎢ ⎥
M=⎢ ⎢ 0 q 0 ··· 0 ⎥
⎥ (24)
⎣ · · · 0 0 ⎦
0 0 ··· q 0
is called the transition matrix. Then, the probability PL (n) that there are no vacant
clusters of size larger than n is found from
PL (n) = (pl )L . (25)
l
One may use a variety of boundary conditions, the simplest of which is perhaps
the periodic conditions which require that the first and the last site of the bundle
to be equivalent, in which case
where tr denotes the trace of the matrix. Thus, all one must do is studying the
eigenvalues of M. Let a1 = p/λ and a2 = q/λ, where λ is the eigenvalue of M.
8.1. Quasi-static Fracture of Fibrous Materials 375
If we define a determinant Dn by
. .
. a1 − 1 a1 ··· a1 a1 .
. .
. a2 −1 0 ··· 0 .
. .
Dn = .. 0 ··· .
. (27)
. ··· ··· a2 −1 0 ..
.
. 0 ··· 0 a2 −1 .
then
Dn = −Dn−1 + (−1)n a1 a2n , (28)
with D0 = a1 − 1. The solution to the recursion relation (28) is
(−1)n Dn = a1 − 1 + a1 a2 + a1 a22 + · · · + a1 a2n = 0. (29)
It is then easy to see that
λn+2 − λn+1 + pq n+1 = 0. (30)
Because M is non-negative, then according to the Perron–Frobenious theorem (see,
for example, Noble and Daniel, 1977) its largest eigenvalue λ is real and unique.
Moreover, it is not difficult to see that λ → 1 as n becomes large. Therefore,
setting λ = 1 − δ, Eq. (30) yields
λ 1 − pq n+1 + O(q 2n ). (31)
and hence for periodic boundary conditions
PL (n) = tr(ML ) = λL
1 + λ2 + · · · + λn λ + O(|λs | ),
L L L L
(32)
where λs is the second largest eigenvalue of M. We thus obtain
PL (n) = [1 − pq n+1 + O(q 2n )]L + O(|λs |L ). (33)
This result agrees with what Duxbury et al. (1986) derived for the electrical break-
down problem discussed in Section 5.2.5. We can now find the failure probability
pf when a stress σ is applied to the bundle by noting that, since failure of the bond
that carries the largest stress causes catastrophic failure, we must have
σf 1
pf = = 1 + n, (34)
σ 2
where σf is the failure stress. Therefore, the probability ps that the fiber bundle
will survive is
L
ps (σ ) = 1 − pq 2σf /σ −1 . (35)
If L and n are large, then Eq. (35) is essentially equivalent to a double exponential
form, also called a Gumbel distribution, a result that was also obtained for electrical
and dielectric breakdown phenomena described in Section 5.2.5.
A more complex situation arises when an intact bond is between two clusters
of vacant bonds, in which case the bond suffers a large stress enhancement. Thus,
for a more complete analysis one must also consider this situation. The same
376 8. Brittle Fracture: The Discrete Approach
technique that was described above can be used to analyze this case, except that
some modifications must be made. For example, the distinct endings that must
be considered are (11), (110), (1100),· · ·,(110· · ·0); (101), (1010), (10100),· · ·,
(10100· · ·0); (1001), (10010), · · ·, and (10· · ·0), each of which occurs with a certain
probability analogous to p(10) , p(100) , and so on. Duxbury and Leath (1994a) then
showed that these more complex configurations do not change the essence of their
analysis described above. After some algebra one obtains
7 8L
PL (n) 1 − [(n + 1)p 2 − pq]q n+1 + O(q 3n/2 ) , (36)
where the second term on the right side of the second equation represents a cor-
rection term for preventing (38) from having unphysical behavior as L becomes
large.
In two other papers, Duxbury and Leath (1994b) and Leath and Duxbury (1994)
developed interesting recursion relations for calculating the failure probability and
average strength of the fiber-bundle model, so that one can numerically study the
behavior of the model [for a different approach, based on calculating the Green
functions, see Zhou and Curtin (1995); for a Green function analysis of fracture
in more general systems see also Zhou et al. (1993)]. As usual, suppose that
{1} denotes an intact (unbroken) bond and {0} a failed one. Then for L = 2 the
surviving configurations are {11, 10, 01}, while for arbitrary L there are 2L − 1
surviving configurations and one failure configuration {0 · · · 00}. The probability
4 (1+n/2)
psn that a bond with n failed neighbors survives is psn = 1 − 0 q(x)dx,
where q(x) is the differential failure probability of a bond. Duxbury and Leath
(1994b) separated the full set of 2L − 1 survival configurations into judiciously
selected subsets. Suppose that a lone surviving fiber is surrounded by failed fibers,
and let {A} be the set of all survival configurations which contain only failed
fibers, and lone fibers, and which are bracketed at both ends by lone fibers. Some
of such configurations are {101, 1001, 10001, 1010, · · ·}. From {A} construct {B},
the set of the configurations one specified end of which must be failed. The failed
configuration at the end can be on the left or the right end, but no distinction is
made between them. A third set {C} is also constructed out of {A} in which both
8.1. Quasi-static Fracture of Fibrous Materials 377
ends of a configuration have failed, e.g., {010, 0100, · · ·}. Finally, suppose that
{P } is the set of configurations with no failed bond, e.g., {1, 11, 111, · · ·}. One
then defines generating functions
∞
∞
∞
A(z) = AL z L , B(z) = BL z L , C(z) = CL z L , (39)
L=3 L=2 L=3
where AL , BL , and CL are the sums, respectively, of the survival probabilities of
the sets {A}, {B}, and {C} for a fixed L. Likewise, a generating function for {P }
is also defined
∞
1
P (z) = (ps0 )L zL = , (40)
1 − ps0 z
L=0
where ps0 is the probability that a bond with no failed neighbors survives. Leath
and Duxbury (1994)
showed that the generating function for the survival configura-
tions, S(z) = L psL zL , is given by (psL is the survival probability for a fixed L)
P (z)[1 + B(z)]2
S(z) = C(z) + . (41)
1 − P (z)A(z)
Since pf L = 1 − psL , where pf L is the failure probability for a fixed L, then
1
f (z) = − S(z) (42)
1−z
where f (z) = L pf L zL , with pf 0 = 0 and psL = 1. We thus obtain
(1 − z)[1 + B(z)]2 − [1 − ps0 z − A(z)]{1 − (1 − z)[f (z) + C(z)]} = 0.
(43)
Expanding identity (43) in powers of zL and setting the coefficient of the zL term
to zero, one finds the following recurrence relation
XL = XL−1 + ps0 DL−1 X − 2DL B − AL + pf 1 AL−1 − B2 BL−2
L−4 (44)
+ (Ai+2 DL−i−2 X − Bi+1 DL−i−1 B),
i=1
in which XL = pf L + CL , and DL Y = YL − YL−1 . Thus one needs AL , BL , and
CL to use recursion relation (44). These are found by defining new subsets {aL,l },
{bL,l }, and {cL,l }, where, e.g., {cL,l } is the set of survival configurations of length
L which end with exactly l failed bonds. Recursion relations are also found for
these new quantities. For example, aL,l = bL−1,l psl , and
L−l−2
bL,l = pf l psl δL−l−1 + bL−l−1,i ps,L+i pf l . (45)
i=1
These recursion relations can then be used efficiently for calculating various quan-
tities of interest. Because of their efficiency, the behavior of the system for large
L, of the order of several thousands, can be studied.
378 8. Brittle Fracture: The Discrete Approach
Figure 8.2. Dependence of the failure probability of the chain-of-bundles model on the
linear size L of the system (after Duxbury and Leath, 1994b).
where 0 < γ (m) < 1 is a parameter that depends on m. Equation (46) has the
general form of a Weibull distribution. Thus, the optimal sample size nmin that
corresponds to the minimum failure probability is obtained from Fn−1 = Fn ,
yielding
nmin ∼ σ −1/γ . (47)
(2) If n nc 1, then the material is in the brittle region, where it is macro-
scopically brittle but microscopically tough. Roughly speaking, the failure of
the material depends on whether the size of the weakest region exceeds nc .
Since, as discussed in Chapter 5, the probability of finding a weak region of
size larger than nc decays exponentially (because in this case the statistics
of the weak or failed regions is described well by percolation statistics), the
cumulative failure probability is of the Gumbel type:
2 3
Fn (σ ) = 1 − exp −an exp(b ln σ/σ m ) , (48)
where a, b and m are fitting parameters. The size dependence of the mean
failure stress σf can then be obtained by neglecting the slow-varying factor
ln σ and taking the median as the average, which then yield
σf ∼ (ln n)−1/m . (49)
(3) For nc ∼ O(1) the material is in the super-brittle regime. This situation arises
when the applied stress is so large that the critical nuclei exist almost ev-
erywhere, and thus almost all the fibers fail simultaneously. The cumulative
failure probability is then simply
Fn (σ ) = 1 − [1 − f (σ )]n , (50)
where f (σ ) is the local strength distribution.
For related work on this problem see Wu and Leath (2000) and Kun et al. (2000).
beams and not at the intersections, and that when the network is deformed, the
angle between the crossing fibers will remain constant. Each fiber-fiber bond has
three degrees of freedom: Horizontal and vertical displacements, and rotations.
An example of a typical realization of such a model is shown in Figure 8.3. Two
distinct cases can be considered. (1) The beams are embedded by a background
material with specific elastic properties, as in, for example, a sheet of paper. (2)
Alternatively, the system consists of a network of the beams alone, as in, for
example, a polymer network.
The elastic properties of the model depend on the aspect ratio w/ lf , as well as
the density p of the fibers, defined as the average total length of fibers in an area
of lf2 . The percolation threshold, or the critical density of the fibers, is given by
pc 5.71lf . (51)
Each fiber contains a segment of length ls which is that part of the fiber that is
between the two intersections that the fiber has with two other fibers. Clearly, the
length of the segments is a random variable, as the fibers are distributed randomly
in the system. The average segment length is given by
ls π pc
= . (52)
lf 11.42(p/pc ) 3.6p
The elastic interaction between two connected bonds is characterized by a stiff-
ness matrix C. If the moment of inertia of the cross section is M = w4 /12, then
8.1. Quasi-static Fracture of Fibrous Materials 381
(53)
The forces acting on the bonds at the segment ends are obtained by multiplying C
by the vector (ux1 , uy1 , ϕ1 , ux2 , uy2 , ϕ2 ), where u = (ux , uy ) is the displacement
vector and ϕ = (ϕ1 , ϕ2 ) is the rotation vector. If ls is short, the bending stiffness
12Yf M/ ls3 = Yf w 4 / ls3 should, as a first approximation, be replaced by the shear
modulus Yf w 2 /[2(1 + νp )ls ], where νp is the Poisson’s ratio of the material.
The fiber network is deformed by, for example, stretching it uniformly in the
x-direction, which means, for example, fixing the edge at x = 0 and pulling in the
positive x-direction the edge which is initially at x = Lx , with the fibers cross-
ing these edges rigidly tied to them. Periodic boundary condition is used in the
y-direction. Computations of the system deformation, when there is a background
matrix, is not straightforward. Typically, a finite-element method, of the type de-
scribed in Section 7.11.3, is used. Several commercial computer programs that
are capable of performing such computations are available. If the system consists
only of the fiber network (with no background material), then the computations
proceed in the same manner that was described for elastic percolation networks
(see Chapter 8 of Volume I). If the fiber density is too low, the system is not rigid
and the elastic stiffness is zero.
To study brittle fracture of the material, a failure criterion must be defined.
Although such criteria will be described in the next section where we discuss more
general discrete models of brittle fracture, we mention a few of them here. One
can, for example, consider a fiber as broken or failed if the axial tension or bending
of its corresponding beam exceeds a pre-set threshold. Alternatively, failure of a
fiber can be defined based on the shear-lag strain, defined as the magnitude of
the jump in the axial strain on a fiber across a bond. A combination of all such
criteria can also be considered, and in fact Åström et al. (1994) studied the case
in which fracture occurred by segment breaking due to axial tension and failure
at a critical value of shear-lag strain. Once the failure criterion is set, the fracture
simulations begin. Each time a fiber fails, the stress and strain distributions in the
network must be recomputed, as the network’s configuration changes dynamically.
As such, the computations are very intensive. Computer simulations indicated that,
the failure of the system at the initial stages occurs more or less randomly, and
thus the fracture process is similar to percolation. Figure 8.4 shows the stress-strain
diagram of the system when relatively few fibers have failed. As expected, up to a
certain strain, the stress-strain relation is perfectly linear which is what is expected
382 8. Brittle Fracture: The Discrete Approach
of brittle materials. Beyond the critical strain, however, the stress shows a generally
downward trend with increasing strains, accompanied by fluctuations that are the
result of having fibers failing at essentially random locations in the system.
However, as the number of the microcracks increases, the facture zone becomes
quasi-1D, populated mainly by such microcracks with no dominating fracture that
can propagate. The absence of a dominating fracture is presumably because of
the random orientations of the fibers that help distribute the applied stress in the
network more evenly than in a regular network where a dominating fracture usually
forms (see below). This behavior is also different from what is usually observed
during fracture of composite, but non-fibrous, materials described in Chapters 6
and 7.
Figure 8.5. Schematics of EMA computation of stress along a fiber (after Åström et al.,
1994).
the strain, it can easily be verified that the following equation due to Cox (1952),
cosh[k( 12 − x/ lf )]
σf (x, k) = Yf x 1 − , (55)
cosh( 12 k)
which he derived by a mean-field approximation, satisfies Eq. (54) and √ the bound-
ary conditions that the stress vanishes at x = 0 and x = lf ; here, k = clf /ls .
The strain x that appears in Eq. (55) is in fact the strain f in the fiber, but use of
x indicates that the strain lies in the x-direction. Note that the average shear-lag
stress is simply ls dσf /dx ∼ kls / lf .
So far, we have assumed that the single fiber embedded in the effective medium is
aligned with the direction of the external strain along which the sheet is stretched
(see Figure 8.5). In general, however, the fibers are distributed randomly, and
therefore one must obtain the orientation dependence of σf . This is, however,
straightforward since, in the absence of transverse Poisson contraction, a rotation
by an infinitesimal field σf yields
σf (x, k, θ) = σf (x, k) cos2 θ, (56)
with θ being the angle of the fiber with respect to the direction of the external strain.
Figure 8.6 compares the predictions of Eqs. (55) and (56) with the simulation results
(Åström et al., 1994) in which k has been treated as an adjustable parameter. It is
clear that the predictions agree well with the simulation results. These simulations
also indicate that k = p(1 + aw/ lf )/pc , where a is a constant.
However, the foregoing treatment is not without problems, especially if it is
further developed in order to predict the elastic stiffness of the network, because
it actually makes the segment stresses correlated along the fibers with reduced
stress close to the fiber ends. A refined treatment of the problem, which we refer
to it as the EMA2, can be developed (Åström et al., 2000) if one combines the
384 8. Brittle Fracture: The Discrete Approach
Figure 8.6. Average distribution of axial stress along fiber for p/pc = 4 and lf /w = 18.8
(after Åström et al., 1994).
probability distribution for the segments’ length with the assumption that the fiber
segments deform only in the energetically most-favorable mode, with the modes
being bending, stretching, and shearing. Since the center of the fibers are distributed
at random in the simulation cell, the probability distribution for their segment
length is known, and is given by
$ %
2p 2p
P (ls ) = exp − ls , (57)
π lf π lf
and therefore the average segment length is, ls = π lf /(2p).
If the deformation of the fiber network is quasi-static, then the fiber segments will
be deformed such that there is force equilibrium at all fiber-fiber bonds, which also
define the global minimum of the total elastic energy of the system. This implies that
the fiber segments will, in general, be deformed in a way that offers the least elastic
resistance. We may define the segments either by bending/shearing or by stretching.
According to the stiffness matrix (53), the bending stiffness modulus is Yf w 4 / ls3 ,
the shear stiffness modulus is Yf w 2 /[2(1 + νp )ls ], while the elongation stiffness
modulus is Yf w 2 / ls . Åström et al. (2000) assumed that a segment deforms only
by bending if the bending modulus is smaller than both the shear ! and elongation
modulus, i.e., if the segment length is such that, ls > lc ≡ w 2(1 + νp ). On the
other hand, if ls < lc , then the segments are assumed to deform by shearing and
stretching. The final ingredient of the model is the assumption that elongation of
a segment is proportional to cos2 θ [similar to Eq. (56)], while bending and shear
are proportional to sin 2θ. We note that the strain field in the effective-medium
treatment does not include any rotation.
8.1. Quasi-static Fracture of Fibrous Materials 385
We can now compute the total elastic energy H of the system which is given by
$ %$ % π/2 $ %
Lx Ly 1 cos4 θ lc 2p Lx Ly
H = px2 Yf w2 dθ exp[−2pl/(πlf )]dl + px2
lf 2 −π/2 π 0 π lf lf
$ % $ %$ %
1 π/2 sin2 (2θ ) lc 2p Lx Ly 1
× Gw2 dθ exp[−2pl/(π lf )]dl + px2 Yf w4
2 −π/2 4π 0 π lf lf 2
∞
π/2 sin2 (2θ ) 2p
× dθ exp[−2pl/(πlf )]dl, (58)
−π/2 4π lc π lf l 2
where G = Yf /[2(1 + νp )]. On the other hand, the elastic energy is related to the
effective stiffness Ce of the network by, H = (1/2)Ce x2 Lx Ly , which means that
the expression for Ce is given by
$ %
pw 2 Yf 2pw 2 e−z 1 −z
Ce = − E1 (z) + 3 + (1 − e ) ,
8lf π lf z 2(1 + νp )
(59)
where z ≡ 2plc /(π lf ), and
∞ e−zx
En (z) = dx. (60)
1 xn
The first test of Eq. (59) is its ability for reproducing the known results in certain
limits. Hence, consider first the limit w → 0. If we rescale the network stiffness,
Ce → Ce /w 2 when w → 0, the fiber network becomes a central-force network,
i.e., a network of simple Hookean springs. Equation (59) then yields Ce ∝ w → 0,
which is expected since the average coordination number of the network is less than
4, and therefore, as explained in Section 8.7.3 of Volume I, the network cannot be
rigid. On the other hand, in the limit p → ∞, which is equivalent to w/ls → ∞,
Eq. (59) predicts that Ce ∝ Yf w 2 p/ lf , implying that in the limit of high p the
network stiffness is simply proportional to Yf multiplied by the density of the
fibers in the network, i.e., the network behaves as an elastic continuum, which is
the expected behavior.
However, there remains one problem to be addressed. In writing down the ex-
pression for the total elastic energy H, Eq. (58), it was assumed that all segments are
deformed. However, below the percolation threshold of the network, Ce = 0, and
no segment is deformed. At, and just above, pc , there are also many segments that
carry no load, while for p pc such segments appear only at the end of the fibers
with a density that can be shown to be about 0.55pc , independent of p (Åström
et al., 1994). Thus, for Eq. (59) to reproduce the correct percolation behavior, one
must make a transformation from p to the density pl of the loaded segments, and
Åström et al. (2000) suggested that p/pc = pl /pc + 0.55 + 0.45/(pl /pc + 1),
which is simply a crossover from p = pc when pl = 0 to pl → p − 0.55pc in
the limits p → ∞ and pl → ∞. Therefore, one should replace the first p on the
386 8. Brittle Fracture: The Discrete Approach
Figure 8.7. Comparison of the predictions of Eq. (62) (curve) with the results of numerical
simulations for w = 0.05 (+), w = 0.06 (×), w = 0.07 (∗), and w = 0.08 (2) (after Åström
et al., 2000).
θf . It can then be shown, using the EMA1 treatment described above, that one
obtains the following expression for the stress σ as a function of the strain x
(Åström et al., 1994):
$ % π/2
2
σ = x w 2 p Yf [cos4 θ + (G/Yf ) sin2 (2θ )]dθ, (64)
π θf (x )
which is obtained from the total elastic energy of the system which, within the
EMA1, is given by
$ % π/2
1 2 2 2
H = x w Yf p Lx Ly [cos4 θ + (G/Yf ) sin2 (2θ )]dθ. (65)
2 π 0
Equation (65) is of course a simplified version of Eq. (58).
As discussed above, as more fibers fail, the fracture zone becomes a narrow,
quasi-1D zone. Thus, in order to create such a zone, one assigns an infinitesimally
lower failure threshold to a band across the network, and then applies Eq. (64)
in this fracture zone which is given a unit width. No fiber fails in the rest of the
network, i.e., Eq. (64) is applied with θf = 0. The result is shown in Figure 8.8 in
which the dashed curve is the equilibrium curve. These predictions are in qualita-
tive agreement with the simulation results shown in Figure 8.4, except that there is
a discontinuity in the predicted stress-strain diagram after the elastic regime (the
regime of a linear relation between σ and x ) ends, whereas the simulations do
not indicate such a sharp and discontinuous change. The disagreement between
the simulation results and the EMA1 predictions becomes progressively stronger
as more fibers fail. The same qualitative trends would have been obtained, had
we used the EMA2 to derive the stress-strain diagram for the fracturing fiber
network. Therefore, although the EMA provides some qualitative insight into the
and Goddard, 1986; Arbabi and Sahimi, 1990b; Sahimi and Arbabi, 1992, 1993,
1996; Sahimi et al., 1993), as well as the Born model described in Chapter 8 of
Volume I (see also below) in which the elastic energy of the system consists of the
contributions by the central forces and a scalar-like term (Hassold and Srolovitz,
1989; Yan et al., 1989; Caldarelli et al., 1994) have all been utilized.
Let us now describe these lattice models. In general, the elastic energy of the
bond-bending (BB) model is given by (Kantor and Webman, 1984)
1 1
H= α [(ui − uj ) · Rij ]2 eij + γ (δθ j ik )2 eij eik , (66)
2 2
ij j ik
where α and γ are the central and BB force constants, respectively, j ik indicates
that the sum is over all triplets in which the bonds j -i and i-k form an angle with
its vertex at i, and eij = 1 if i and j are connected, and eij = 0 otherwise. The first
term on the right side of Eq. (66) represents the contribution of the stretching forces,
while the second term is due to BB forces. The precise form of δθ j ik depends on
the microscopic details of the model. In the most general form, if bending of all
pairs of bonds that have one site in common, including the collinear bonds, is
allowed, then (Arbabi and Sahimi, 1990a)
(uij × Rij − uik × Rik ) · (Rij × Rik )/|Rij × Rik |, Rij not parallel to Rik ,
δθ j ik =
|(uij + uik ) × Rij |, Rij parallel to Rik ,
(67)
where, uij = ui − uj . For all 2D systems, Eq. (67) is simplified to
where Rij is the unit vector along the line from i to j , and α1 and α2 repre-
sent, more or less, two adjustable parameters. The first term of Eq. (68) is the
energy of a network of central-force springs, i.e., Hookean springs that trans-
mit force only in the Rij direction, but do not transmit shear forces, whereas
the second term is a contribution analogous to scalar transport (for example,
the power dissipated in conduction), since (ui − uj )2 represents the magnitude
of the displacement difference ui − uj . The Born model can be derived from
linear continuum mechanics by discretizing the linear equation that governs the
elastic equilibrium of a solid, i.e., ∇ · σ = 0 (where σ is the stress tensor), and
using the usual relation, σ = λ(∇ · u)U + µ[∇u + (∇u)T ], where λ and µ are
the usual Lamé constants, and U is the identity tensor (see Section 8.4 of Vol-
ume I for details). If this is done, then one obtains, α1 = 2(1 − νp )/(1 + νp ), and
8.2. Quasi-static Fracture of Heterogeneous Materials 391
α2 = 2(1 − 3νp )/[4(1 − νp )], where νp is the Poisson’s ratio. However, in this
form, the elastic energy given by Eq. (68) will not be rotationally invariant, thus
violating a fundamental physical requirement for an elastic energy representation
of a solid material. Therefore, Eq. (68), in which α1 and α2 are treated as adjustable
parameters, is a semi-empirical representation of materials.
The Born model may be considered as an analogue of a 3D solid in plane-stress
with holes normal to the x-y plane, or as a 2D solid with the Poisson’s ratio defined
as the negative of ratio of the strain in the y-direction to that in the x-direction,
when a stress is applied in the x-direction but none is applied in the y-direction.
Results for a 3D solid in plane-strain can be generated from those of this model
using the transformation νp = νp /(1 + νp ), where νp is the Poisson’s ratio for the
plain strain.
Let us mention another interesting way of generating a BB model. In their studies
of brittle fracture, Chung et al. (2001) generated a spring network by molecular
dynamics simulation, starting with a random distribution of spheres that interact
with each other through certain potentials. The system would then be allowed to
reach equilibrium, after which the centers of the spheres that were not separated by
a distance larger than a certain limit were connected by springs. Both the central
and BB forces were included in the network so obtained.
The spring lattices are suitable models for simulating a fracture process in mate-
rials that are under shear or tension. However, one should use the beam model (see
Chapter 8 of Volume I for more details) (Herrmann et al., 1989a; de Arcangelis et
al., 1989; Tzschichholz, 1992,1995; Tzschichholz et al., 1994; Tzschichholz and
Herrmann, 1996) when external compressional forces are imposed on the system,
since a spring cannot break under compression. In the beam model, in addition to
the central and BB or angle-changing forces, torsional forces also contribute to the
elastic energy H of the lattice. We believe, however, that, except when external
compressional forces are imposed on the system, the BB model is a completely
realistic representation of the elastic energy of disordered materials. Recall that, as
discussed in Chapters 8 and 9 of Volume I, the BB model is capable of describing
the elastic properties of polymers, glasses, ceramics and powders, and hence use
of more complex models for the elastic energy of the material is not necessary.
In addition to the above models, a model based on discretization of the following
equation (sometimes referred to as the Lamé equation)
(λ + µ)∇(∇u) + µ∇ 2 u = 0, (69)
where λ and µ are the usual Lamé constants, has also been used (Herrmann et al.,
1989b).
Sahimi and Goddard (1986) suggested that three general classes of disorder may
be incorporated into such model, which are as follows.
(1) Deletion or suppression of a fraction of the bonds either at random or in a
prescribed fashion, so that the material’s heterogeneity is of percolation-type.
The suppressed or deleted bonds may, for example, represent the microporosity
or some type of defect in the system before the fracture process began.
392 8. Brittle Fracture: The Discrete Approach
(2) Random or correlated distribution of the elastic constants eij of the bonds.
The idea is that in real heterogeneous materials the shapes and sizes of the
elastic zones through which stress transport takes place may be statistically
distributed, resulting in a different eij for each zone, or bond in the lattice
model, that follows some type of a statistical distribution. Such a model may
be appropriate for a composite material that, for example, consists of several
constituents, each of which has its own elastic properties.
(3) Random or correlated distribution of the critical thresholds at which the lin-
ear constitutive relation that describes the stress-strain relation in the beam
or spring breaks down. For example, in shear or tension each bond may be
characterized by a critical length lc , such that if it is stretched beyond lc , it
breaks irreversibly. Such a threshold can be estimated experimentally by eval-
uating macro tensile strength of the material. Alternatively, each bond can be
characterized by a critical force (stress) Fc (σc ), such that if it suffers a force
(stress) larger than Fc (σc ), it breaks irreversibly. Under compression, a beam
breaks if it is bent too much. The idea for using this type of disorder is that
a solid material made up intrinsically of the same material (the same elastic
constant eij everywhere) may contain regions having different resistances to
breakage under an imposed external stress or potential due to, for example, the
presence of defects during its manufacturing or formation process. Depending
on the intended application, we may use any combination of the three types
of disorder. For example, one may model the disordered material with fractal
lattices with bonds that have statistically-distributed properties (such as their
elastic constant). Because of their fractality, such models have low connectiv-
ities and large porosities, and may be relevant to transgranular stress corrosion
cracking of ductile metal alloys, such as stainless steel and brass (Sieradzki
and Newman, 1985). They may also be relevant to stress and crack propaga-
tion in weakly-connected granular media, such as sedimentary rocks. We do
not, however, consider them here as they have not been studied extensively.
Another important source of disorder in stressed materials is the so-called resid-
ual stress variations, which are caused by, among other things, thermal expansion
mismatch. The appropriate elastic lattice models with bond mismatch were de-
scribed in Section 9.7 of Volume I. We will not discuss the effect of this type of
disorder on fracture, although they can be analyzed by modification of the models
that are described here (see, for example, Curtin and Scher, 1990a,b; Sridhar et
al., 1994).
After selecting the lattice and the form of the elastic energy of the system (i.e.,
the types of forces that are operative in the lattice), we specify the type of the
heterogeneity that the material contains. If the disorder is of percolation-type (type-
1 heterogeneity described above), then its inclusion in the lattice is straightforward
and needs no discussion. For types-2 and 3 heterogeneities described above, their
statistical distribution must be specified. A statistical distribution that has been
used widely is
f (x) = (1 − ζ )x −ζ , (70)
8.2. Quasi-static Fracture of Heterogeneous Materials 393
where x is any property of the lattice that is statistically distributed and repre-
sents its heterogeneity, and 0 ≤ ζ < 1. The advantage of the distribution (70) is
that, varying ζ allows one to generate distributions that are very narrow (ζ → 0)
or very broad (ζ → 1), and therefore one can study the extent to which such ex-
treme distributions affect failure phenomena. Note that ζ = 0 represents a uniform
distribution, while Roux et al. (1988) showed that, in the limit ζ → 1, fracture be-
comes equivalent to a type of percolation. A great advantage of the lattice models is
that, any type of statistical distribution f (x) of the heterogeneities can be used. For
example, de Arcangelis et al. (1989) used, in addition to (67), a Weibull distribution
2 3
f (x) = mλ−m x m−1 exp −(x/λ)m , (71)
where 2 ≤ m ≤ 10 supposedly describes many real materials.
After specifying the lattice type, the form of the elastic energy H, and the type
of disorder, the boundary conditions must be specified. One can, for example,
use shear, uniaxial tension or compression, uniform dilation (i.e., pulling a lattice
equally in all directions), or surface cracking which is used for simulating fracture
of a thin film of a material attached to a substrate (for example, thin polymeric
coatings, or paints, or even mud). In this case, each site of the lattice is connected
by a spring to the substrate which has a lattice constant larger than the original
lattice. In this way all the bonds are equally overstretched without having applied
any force on a boundary of the lattice, implying that no external boundary is in
fact needed, and one can use periodic boundary conditions in all directions.
The simulations can now begin. One must compute the distribution of the nodal
displacements (and rotations, if such motions are allowed), from which the forces
(and stresses) exerted on all the bonds are computed. The procedure for doing so
consists of minimizing the total elastic energy of the system with respect to the
displacements of the internal nodes of the lattice (and their rotations, if such motion
is allowed). Because of the assumption of brittleness, these equations are linear
and therefore, subject to the boundary conditions imposed on the system, can be
solved by one of several methods that are available for solving such equations. If
very high precision is needed, then the conjugate-gradient method (see Chapter 9
for a description of this method) is the best technique to use.
After computing the initial distribution of the stresses (and strains) in the lattice,
a criterion for nucleation of the microcracks must be specified. The criterion,
however, depends on the type of material that is being studied. For example, if
each elastic bond is a rubber band, then it will tear apart when stretched beyond
a certain limit. Thus, for example, we assign a threshold lc for the length of the
bonds, which is selected from the probability density functions described above.
Then, in terms of lc , the breaking criterion is that a bond breaks if its length in
the deformed lattice exceeds its lc . Alternatively, among all the bonds that have
exceeded their lc , the one with the largest deviation from its lc breaks first. The
idea is that in a deformed material, the weakest point of the system fails first.
However, if the elastic bond represents, for example, a glass rod, then it will
break if it is bent too much. One must of course use a lattice of beams for modeling
such a material. Therefore, a good strategy would be devising a breaking criterion
394 8. Brittle Fracture: The Discrete Approach
that is a combination of both stretching and bending. For example, in the beam
model one can use the criterion (de Arcangelis et al., 1989) that the beam with
largest value of the following quantity
$ %2
F |Mi | |Mj |
pb = + max , , (72)
Fc Mc Mc
breaks, where F is the longitudinal force acting along the beam and Fc its critical
value, and Mi and Mj are the moments applied on the two adjacent sites i and
j of the beam, with Mc being the critical threshold of the moment. Both Fc and
Mc are distributed according to some probability density functions and can, in
fact, represent the heterogeneity of the material (type-3 disorder discussed above).
The first term of (72) describes breaking due to stretching, while the second term
is representative of breaking due to bending. One can distribute, for example, Mc
over a broader range than Fc (and vice versa) as a measure of susceptibility of the
material to breaking by bending as opposed to stretching.
Other failure criteria of this type can be, and have been, used. For example,
a combination of Fc and lc can also be used for setting up a breaking criterion
(Arbabi and Sahimi, 1990b): one breaks that bond for which the ratio R = lm l/ lc
is maximum, where l is the current length of the bond in the deformed lattice and
lm is the maximum length that any bond in the lattice has, or break the bond for
which U = F lc /Fm is minimum, where F is the total force that the bond suffers,
and Fm is the maximum force on any bond in the lattice. Sridhar et al. (1994) used
the criterion that a bond breaks if its strain energy exceeds a critical value. The
flexibility that the lattice models afford one in using almost any type of failure
criterion is one important advantage of such models.
Another advantage of such lattice models is that, depending on the intended
application, any failure criterion can be used without imposing any undue difficulty
on the computations. For example, often portions of a material are damaged but
do not break during the fracture process. The damaged area can be modeled by
lowering the breaking threshold of the bonds that are in the vicinity of the growing
crack. It is clear that the bonds that are damaged in this way are more likely to
break in the next step of the simulations than the undamaged, unbroken bonds.
One can also model short-lived damage by considering at step n of the simulations
the quantity
pb = pb (n) + β0 pb (n − 1), (73)
where pb (n) is the usual quantity that one uses for breaking criterion at step n
without considering the damage [for example, Eq. (72)], and 0 < β0 < 1. Then,
the breaking criterion must be based on pb , and pb (n − 1) indicates the effect of
“memory” of the damage that occurred during the previous step of the simulations.
In some fracture processes, such as stress corrosion, the damage accumulates
during the entire process, and therefore the breaking criterion must somehow
reflect this. A simple algorithm for taking this effect into account is as follows
(Herrmann et al., 1989b). One assigns a counter c(n) to each bond of the lattice
that is susceptible to damage, e.g., those in the “vicinity” of the growing crack,
8.2. Quasi-static Fracture of Heterogeneous Materials 395
where the boundaries of this region must also be specified. For example, one can
consider all the unbroken bonds that are up to a certain distance from the cracked
area. At the beginning of the simulations, all the counters are set to zero, c(0) = 0,
for all the bonds that may be damaged. Then, at each step n of the simulations, one
calculates pb (n), the quantity based on which the breaking criterion is applied,
and computes
1 − c(n − 1)
α(n) = . (74)
pb (n)
The bond with the smallest value of α, namely αm , is then broken. Then, all the
counters are updated by
c(n) = αm pb (n) + βc(n − 1), (75)
so that c(n) “accumulates” the damages from all the previous steps of the process.
Clearly, if β = 1, then the damage is irreversible. The limit β → 0 corresponds to
criterion (73) with β0 = 1.
The failure process is then initiated by selecting the bond (or bonds) that must
break. Two different “dynamics” of fracture propagation can be studied. In model
1 only one bond is broken at each stage of the simulation, which is equivalent
to assuming that the rate at which the elastic forces relax throughout the network
is much faster than the breaking of one bond. In model 2 (and depending on the
failure criterion), all bonds that meet the failure criterion are broken. Most of the
studies so far have used model 1, and the properties of model 2 have not been
studied extensively. Breaking a bond is equivalent to removing it from the lattice
(by, for example, setting its elastic constant to zero), after which one recalculates
the stress and strain distributions for the new configuration of the lattice, select the
next bond(s) to break, and so on. If the external stress or strain imposed on the lattice
is not large enough to break any bond, it is gradually increased. The simulation
continues until a sample-spanning crack is formed. As mentioned above, instead of
removing a failed bond from the lattice, one may reduce its elastic constant. In this
case, one observes the interesting phenomenon of crack arrest (Li and Duxbury,
1988).
very strong and very weak regions. For the growing crack to take advantage of the
weak regions, it must find its path which could be quite tortuous, so much so that it
may result in fragmentation of the material (Åström and Timonen, 1997a). We will
discuss the phenomenon of fragmentation later in this chapter. On the other hand,
when weak disorder is present in a solid, a catastrophic crack is formed quickly
that spans the system, and therefore the mechanical failure of the system is very
fast.
The number and shape of the cracks also depend on the dimensionality of the
system. In 3D stress enhancement at the tip of the growing cracks is much weaker
than in 2D, so that even mild disorder can give rise to very complex fracture
pattern in 3D. Moreover, the type of the boundary conditions applied to the system
is also important. If a stress (instead of a strain) is applied to the material, then
the system would fail very quickly soon after the first microcrack is formed, even
if the stress is applied slowly to the system. Therefore, similar to real fracture
tests, there are significant differences between a stress-controlled and a strain-
controlled fracture test in such models. Figure 8.9 shows three different stages of
fracture of a 2D triangular lattice. The bonds of the lattice suffer only stretching
(central) forces, disorder was generated in the lattice by removing 10% of the
bonds at random before deformation of the lattice began, and the deformation was
caused by applying a strain to one face of the lattice, while the opposite face was
fixed. The top panel shows the initial configuration of the lattice before it was
deformed. The middle panel represents the system when it loses its rigidity. Recall
from Chapter 8 of Volume I that if only central forces act on the bonds, the lattice
loses its rigidity at a fraction of uncut springs which is significantly larger than the
connectivity threshold. For example, the connectivity threshold of the triangular
lattice is at pc = 0.347, while a triangular lattice of Hookean springs loses its
rigidity at p 0.641. However, the rotational freedom of the bonds (which do
not contribute to the total elastic energy of the system) is not lost yet. At this
point, the lattice contains a strip of bonds with zero shear modulus, which can be
thought of as a type of shear band. Note that the shear band can move at most
one bond before the shear modulus is re-established. Final failure of the lattice,
which occurs due to formation of a macroscopic crack that splits the system into
two pieces, occurs very close to the shear band. In contrast with Figure 8.9, Figure
8.10 presents an example of the microcracks that are produced in the same lattice
in which both the central and bond-bending (angle-changing) forces contribute
to the elastic energy of the system, and the distribution of the heterogeneities is
broad, following Eq. (70).
Other interesting crack shapes can also be produced by such models. For exam-
ple, Herrmann et al. (1989b) considered a square lattice of beams with periodic
conditions in the horizontal direction, while the top and bottom of the lattice was
sheared. They defined a probability pb slightly more general than Eq. (69), given by
pb = [F 2 + q · max(|Mi |, |Mj |)]η , (76)
where F is the traction (and/or compression) force applied on a beam, and q and
η are two free parameters, with q being a measure of the affinity of a beam to
8.2. Quasi-static Fracture of Heterogeneous Materials 397
breaking by bending, and η having no apparent physical meaning. A beam was re-
moved at the center of the lattice in order for microcracking to be initiated. Figure
8.11 presents three fracture patterns. The one at the top was produced by using a
breaking criterion based on Eqs. (74) and (75), where pb was assumed to be given
398 8. Brittle Fracture: The Discrete Approach
by Eq. (76), and q = 0, β = 1 and η = 1, while the middle one was generated
based on Eq. (73) with β0 = 1 and η = 0.2. Evidently, decreasing η increases the
tendency of the growing fracture to branch out. Also shown at the bottom of Figure
8.11 is the fracture pattern in an alloy, Ti11.5 Mo6 Zr4.5 Sn, aged 100 hours at 750 K
and cracked under increasing stress intensity. There are certain similarity between
the experimental pattern and what is shown in the middle of the figure.
Let us point out that, under certain conditions, the set of the broken bonds (the
microcracks) can form a fractal network. The fractality of this set, and that of the
sample-spanning fracture, will be discussed later in this chapter.
Figure 8.12. Young’s modulus of the central-force (triangles) and the bond-bending (2)
models (with γ /α = 0.04) versus the fraction p of unbroken bonds. Also shown are the
modulus of a simple-cubic lattice in bond percolation (circles), as well as the experimental
data (dashed curve) (after Arbabi and Sahimi, 1990b).
within the range of the experimental data. Moreover, except for p 1, the re-
sults produced by the central-force model do not agree with the experimental data
because such systems fail at high values of p. For p ≥ 0.5, the results predicted
by the random percolation model do not agree with the data as well as the lattice
model of fracture, presumably because the percolation threshold of the lattice is
low (pcB 0.25) and, as a result, the predicted modulus is somewhat large.
An interesting study by Curtin (1997) demonstrated the effect of disorder on
quasi-static fracture. He investigated the fracture toughness of heterogeneous ma-
terials using a simple-cubic lattice of springs with distributed toughness. He found
that the overall toughness of the lattice or, equivalently, the stress σf to initiate the
first microcrack, is a random variable that depends on the width of the toughness
distribution of the individual springs. This by itself is not surprising (see below).
What was interesting was the finding that, for narrow distributions, the toughness
was found to be controlled by the nucleation of the kinks at the weakest springs,
whereas for broad distributions the toughness was controlled by the highly rigid
regions of the system that pin the growth of the fracture front. However, the dif-
ference between the toughnesses of materials with narrow and broad distributions
400 8. Brittle Fracture: The Discrete Approach
was found to be small, hence suggesting that simple disorder alone (such as a
narrow and uniform distribution of the thresholds) cannot solely be responsible
for the variety of fracture behavior seen in experiments, and more complex factors
must play an important role. We will come back to this issue later in this chapter.
1 1
Tf (f + dν + x) = (fsc + dν), (78)
2 2
where fsc = f + x is the elasticity exponent of the Swiss-cheese model, and
x = 3/2 and 5/2 for d = 2 and 3, respectively. Thus, for a 2D material, one obtains
Tf 4.06, in good agreement with the experimental measurement of Benguigui
et al. (1987) who also measured the fracture strength of a 2D Swiss-cheese model
(by punching holes into their material) and reported that, Tf 4.0 ± 0.1.
Rigorous upper and lower bounds for Tf have also been derived (Ray and
Chakrabarti, 1985a,1988). In a lattice near the percolation threshold, representing
a composite material with strong disorder, the nearest-neighbor uncut bonds or sites
form a type of “super-lattice”—a very large cluster made of tortuous links (long
strands made of many bonds)—that cross each other at nodes that are separated
by an average distance ξp , the correlation length of percolation. Therefore, in a
d-dimensional system, the external stress σ is shared by ξpd−1 number of parallel
402 8. Brittle Fracture: The Discrete Approach
links, implying that the stress per link is σl σ ξpd−1 . Then, the total strain is,
f/ν
= σ/Y σ ξp , where we have used the power-law dependence of the Young’s
modulus on p near pc . The total strain is shared by ξp−1 number of links, which
1+f/ν
means that the strain per link is given by, l σ ξp . Therefore, the elastic
energy per link is given by
d+f/ν
Hl σl l σ 2 ξp . (79)
We now recall that only the bonds in the backbone of the lattice, i.e., the multiply-
connected bonds of the sample-spanning cluster, contribute to stress and/or strain
transfer in the system, the total number M of which is, M ∝ ξpDbb , where Dbb is the
fractal dimension of the backbone. If Hl is shared equally by all such bonds, then
d−D +f/ν
the energy per bond Hl /M is of the order of σ 2 ξp bb . However, because
of the assumption of equal sharing by all the M bonds of the links, this value of
Hl /M underestimates the strain energy per bond, since many of the bonds either
do not support any stress at all or support very small stresses, because the backbone
of a percolation lattice is multiply connected. Thus, if we assume a fixed energy
threshold for breaking each bond, lattice fracture will occur for σf = σ for which
d−D +f/ν
Hl /M ∼ σ 2 ξp bb exceeds the threshold. Therefore, if σf follows Eq. (77),
then, because we underestimate Hl /M, we must have
1
Tf ≥ [f + (d − Dbb )ν]. (80)
2
On the other hand, recall that the number of singly-connected (red) bonds, i.e.,
1/ν
those that, if cut, would split the backbone into two pieces, is Mr ∼ ξp . If all the
strain energy is supported by such bonds, then one obtains an overestimate of the
d+f/ν−1/ν
elastic energy per bond, Hl /Mr ∼ σ 2 ξp . Using the same argument that
we utilized for the lower bound, we obtain
1 1
[f + (d − Dbb )ν] ≤ Tf ≤ (f + dν − 1). (81)
2 2
Using the numerical values of the various exponents, f 3.96 and 3.75, and
Dbb 1.675 and 1.87, for 2D and 3D systems, respectively, we obtain, 2.22 ≤
Tf ≤ 2.81 for d = 2, and 2.37 ≤ Tf ≤ 2.7 in d = 3. These bounds are perfectly
consistent with the experimental and numerical estimates of Tf given above.
Somewhat sharper bounds were proposed by Bergman (1986) who suggested
that
f − νDmin ≤ Tf ≤ f − 1, (82)
where Dmin is the fractal dimension of the shortest paths, or the chemical paths, on
the backbone of percolation clusters. These bounds, together with the estimates,
Dmin 1.13 and 1.34 for 2D and 3D systems, respectively, imply that
2.45 ≤ Tf ≤ 2.96, d = 2, (83)
2.58 ≤ Tf ≤ 2.76, d = 3, (84)
8.2. Quasi-static Fracture of Heterogeneous Materials 403
which agree nicely with the simulation results. Moreover, in both 2D and 3D the
estimated Tf is close to the lower bound (82), and therefore the relation Tf =
f − νDmin cannot be ruled out.
where ρ(σ ) is the density of cracks weaker than the stress σ or, equivalently, the
density of the cracks that will start propagating at and above the stress σ . Equation
(87) is due to the fact that the sample material survives if each of the cracks within
it survives. We now show that the WD and GD correspond to two limiting cases
of Eq. (87).
Consider first the derivation of the WD. This distribution arises if ρ(σ ) is given
by a power law in σ , which will be the case if the density ρ(l) of the linear cracks
is of power-law type. This type of distribution does not arise in materials with a
random distribution of vacancies or voids, unless the material is precisely at its
404 8. Brittle Fracture: The Discrete Approach
distributions can be made if we rewrite Eqs. (85) and (86) in alternative forms. If
we define a quantity A by
ln[1 − FL (σf )]
A = − ln − , (92)
Ld
then the WD can be rewritten as
$ %
1
Aw = a1 ln + b1 , (93)
σf
while the GD is rearranged as
1
AG = a2 + b2 . (94)
σfδ
These two equations predict linear variations of Aw with ln(1/σf ) and of AG with
1/σfδ . The exact value of δ has not been determined, but in general 1 ≤ δ ≤ 2.
The conditions under which Eq. (85) or (86) may provide accurate representation
of fracture strength data for a given material are not completely clear yet. It appears
(Sahimi and Arbabi, 1993b) that in highly heterogeneous solids neither equation
is very accurate, although the WD appears to perform better. On the other hand,
in weakly-disordered materials, or those far from the percolation threshold (i.e., a
material with few vacancies or voids), the GD may be a better representation of
the distribution of fracture strengths. For example, Figure 8.13 presents the fit of
the fracture simulation results to a GD with δ = 1 for a triangular network with the
central and bond-bending forces, in which before deformation and fracture of the
lattice started, 10% of the bonds had been removed at random. Simulations with
the central force (Beale and Srolovitz, 1988) and the Born models (Hassold and
Figure 8.13. Fit of the simulation results for fracture of a triangular network with stretching
and bond-bending forces (with γ /α = 0.1) to the Gumbel distribution. Before fracturing,
10% of the bonds were removed at random (after Sahimi and Arbabi, 1993).
406 8. Brittle Fracture: The Discrete Approach
Srolovitz, 1989) also seem to indicate that the GD is accurate if the system is far
from pc , although Hassold and Srolovitz (1989) reported equally accurate fit of the
data with the WD, and Hansen et al. (1989) also reported some deviations from the
GD in their central-force model. Curtin and Scher (1992) discussed the conditions
under which a WD may be appropriate for representing the distribution of fracture
strength. We note, however, that as the percolation threshold is approached, neither
distribution seems to be very accurate.
Fracture behavior of a material at its percolation threshold pc , or equivalently
above its pc but at length scales L ξp , depends on the broadness of the distri-
bution of its heterogeneities, and deserves to be discussed (Sornette, 1988). In a
percolation system far from pc , there are many multiply-connected paths, called
macro-links, which support stress transport. In such a system, the distribution of
fracture strength is the result of one or both of the following factors:
(1) Fluctuations of the individual characteristics of the microscopic regions (for
example, their breaking threshold lc and/or their elastic constants) of the
material.
(2) Fluctuations of the macro-link sizes L around the percolation correlation
length ξp . If the characteristics of the microscopic regions are all the same
(i.e., if the material is not made of different constituents with different proper-
ties), the first factor cannot contribute to the distribution of fracture strength.
As pc is approached, two changes take place: First, one has fewer macro-links
and, secondly, compared with those of the long macro-links, the contribu-
tions of the shorter macro-links to stress transport become negligible. Thus,
macro-link to macro-link fluctuations also decrease. At pc , there is only one
huge macro-link, and therefore all the fluctuations disappear completely and
the distribution of fracture strength must be a delta function. However, if,
for example, the elastic constants of the microscopic regions of the material
are statistically-distributed quantities (which is often the case in real hetero-
geneous solids), then region-to-region fluctuations exist and the distribution
of fracture strength is a meaningful quantity to define, calculate, or measure.
This is particularly important for disordered materials modeled by continuum
percolation which usually possess a broad distribution of the elastic constants
of the channels through which stress transport takes place. The distribution of
fracture strength in such materials is a Weibull-like distribution, rather than the
GD. This is supported by the lattice simulations of Sahimi and Arbabi (1993).
c
σfδ = , (95)
A(p) + B(p) ln L
where A(p) and B(p) are simple functions of p, and c is a constant. In particular,
B(p) is small for p close to unity, and B(p) → ∞ as p → pc . Simulations with
the central-force model (Beale and Srolovitz, 1988), the Born model (Hassold
and Srolovitz, 1989), and the central-force and bond-bending model (Sahimi and
Arbabi, 1993) all support the accuracy of Eq. (95).
In general, one is also interested in the scaling behavior of the external stress σ
or force F for breaking the lattice and its dependence on its linear size, since in
practice not only can this quantity be measured easily, but it also provides us with
important insight into a material’s structure (that is, whether the material is strong
and difficult to break versus being weak and easily breakable, both of which depend
on its morphology and size). Since this force is, for example, proportional to Y ,
where is a displacement or strain, a plot of the stress σ , or the force F , versus
can be compared with the traditional stress-strain diagrams that are measured
for composite solids. However, changing the parameters of the lattice model will
result in a wide variety of such diagrams. Therefore, instead of presenting the
results for each model and lattice size separately (which would be impossible), we
may try to collapse the data for all values of L, the linear size of the lattice, onto a
single curve. If the data collapse is possible, then such diagrams possess universal
properties which could be exploited for practical purposes.
Figure 8.14 presents (Sahimi and Arbabi, 1993) the results for the triangular
lattice with sizes L = 50 and 70, in which both the central and bond-bending
forces act on the springs. The thresholds were distributed according to Eq. (70) with
ζ = 0 (a uniform distribution). However, as can be seen, the data collapse is not
complete and three distinct regimes can be discerned. The first regime represents
the initial stages of crack growth and is far from the maximum of the curve. In
this regime microcracking propagates at a relatively slow rate and is more or
less similar to a percolation process, as the springs break essentially at random. As
microcracking proceeds, one eventually arrives in the second regime in the vicinity
of the maximum in which microcracking is intense and the lattice is relatively close
to its macroscopic failure point. Beyond the maximum, the system is in the so-
called post-failure regime, and is highly sensitive (unstable) to small variations
in the applied stress or strain. The general features and shapes of these curves
are in good qualitative agreement with direct experimental measurements and
observations for brittle fracture in various types of disordered solids.
408 8. Brittle Fracture: The Discrete Approach
Figure 8.14. Collapse of stress-strain data for fracture of a triangular network with
stretching and bond-bending forces (after Sahimi and Arbabi, 1993).
and Sahimi, 1990b; Sahimi and Arbabi, 1993), 1 2 ± 0.1, and ψ 0.2, which,
together with the results for 2D lattices, suggest that for a d-dimensional system
Ld−1
F ∼ φ(/Ld−1 ), (98)
(ln L)ψ
where 0 ≤ ψ ≤ 0.2. Our unpublished simulation results for quasi-static fracture
of a BCC lattice with central forces and lattice sizes of up to L = 32 indicated
that the results do follow Eq. (98). Note that Eq. (98) has a simple interpretation:
Ld−1 is the surface area on which the external force is applied, and (ln L)ψ is the
manifestation of the sample-size effect on the fracture process.
One can also study the variations of F with Nb , the number of bonds that break
during fracture. We assume that
F ∼ L3 φ(Nb /LDf ). (99)
An equation similar to (99) was used in the electrical breakdown problem; see
Chapter 5. Equation (99) implies that if the force F is plotted versus Nb /LDf , then
the results for various lattice sizes and parameters of the model should collapse onto
each other. Figure 8.15 presents such a data collapse (Sahimi and Arbabi, 1993). de
Arcangelis et al. (1989) found for their 2D models that, 3 0.75 and Df 1.7,
consistent with their results obtained with Eq. (96). On the other hand, Arbabi
and Sahimi (1990b) and Sahimi and Arbabi (1993) found that 3 1 ± 0.05, and
Df 1.7 ± 0.1 in 2D, and 3 2 ± 0.1, and Df 2.3 ± 0.2 in 3D, consistent
with their results and Eqs. (97) and (98). Note that Df represents the fractal
dimension of the set of all the broken bonds.
As discussed by de Arcangelis et al. (1989), there are other interesting scaling
features of these models. For example, one can study how the stress at the maximum
of the diagram of the type shown in Figure 8.14 scales with the linear size L of
Figure 8.15. Collapse of the data according to Eq. (99) (after Sahimi and Arbabi, 1993).
410 8. Brittle Fracture: The Discrete Approach
the sample. Similarly, one can look at the scaling of the number of the broken
bonds Nb that corresponds to this stage of the fracture at the maximum. Finally,
the number of broken bonds at the end of the fracture process also scales with L
as LDf , with Df 1.7 in 2D.
Figure 8.16. Logarithmic plot of the experimental data for fracture load (open circles) and
yield load (filled circles) versus p − pc (after Benguigui et al., 1987).
found to be more accurate. Figure 8.17 presents their data and their fit to the GD.
Evidently, the pre-fracture porosity of the material used in these experiments was
very low, so that the porous ceramic was far above its percolation threshold.
In another set of interesting experiments, Li and Sieradzki (1992) studied me-
chanical breakdown of random porous Au, a new material specifically designed for
their experiments. They used digital image analysis to characterize the microstruc-
tures of their samples which varied by more than two orders of magnitude in length
scale. The porous Au underwent a microstructurally controlled ductile-brittle tran-
sition. Such a transition had already been predicted by Sahimi and Goddard (1986),
based on the broadness of the heterogeneity distribution in their lattice model of
fracture, and in the numerical simulations of Kahng et al. (1988) using the scalar
(fuse) model fracture processes (see Chapter 5). These results provide support for
usefulness of the lattice models for describing quasi-static fracture processes in
composite solids. Other relevant experimental data are discussed below.
Figure 8.17. Fit of experimental data for fracture strength of porous silicon extrudates to
the Gumbel distribution (after van den Born et al., 1991).
There are at least two ways of comparing a fractured lattice with a percolating
one. The first method is based on comparing the force distributions (FDs), i.e., the
distributions of the forces that are exerted on the bonds of the lattice, and their
moments in the two systems. We already described in Chapter 8 of Volume I (see
Sections 8.6.3 and 8.11.4) the FD of elastic percolation networks, first computed by
Sahimi and Arbabi (1989). As discussed by Sahimi and Arbabi (1993), the initial
stages of brittle fracture and percolation processes in a lattice are more or less
similar. That is, during the initial stages of fracture growth, the bonds that break
are distributed essentially at random in the lattice, unless the lattice is uniform, or
its disorder is very weak. In these initial stages, the stress enhancement at the tip
of a given microcrack is not strong enough to ensure that the next bond that breaks
would be at the tip of the present microcrack. However, as more microcracks
nucleate the effect of stress enhancement becomes stronger, and deviations from
random percolation increase. Beyond a certain point in the growth of the cracks,
there will be no similarity between the two processes. Hence, one is naturally
interested to locate the point at which a fracture process starts to deviate from
a percolation phenomenon. The key clue is already provided in the stress-strain
diagrams discussed above. Equations (96)–(98) are manifestation of finite-size
scaling which represent the fracture data for various lattice sizes up to the maximum
of the stress, beyond which it breaks down. This type of finite-size scaling is also
valid for percolation networks for any p (the fraction of the intact bonds) in the
interval pc ≤ p ≤ 1 (so long as the linear size L of the network, L < ξp , where ξp
is the correlation length of percolation), albeit with different exponents and scaling
functions. Therefore, in the type of disordered lattices that we are considering here,
fracture and percolation are more or less similar up to the maximum in the stress-
strain diagram of the fractured system, i.e., in the regime in which finite-size scaling
is applicable to the fracture data, but they are not similar beyond this point.
Similar to elastic percolation networks analyzed in Chapter 8 of Volume I, one
can also calculate the moments M(q) of the FD in a fractured lattice and study
their scaling with the lattice’s linear size. Thus, one writes
q
M(q) = nFi Fi , (100)
i
where nFi is the number of bonds that suffer a force with a magnitude Fi . Similar
to elastic percolation networks, each moment of the FD scales with the linear size
L of the fractured lattice as
Herrmann et al. (1989a) and de Arcangelis et al. (1989) calculated the correspond-
ing exponents τ̃ (q) for 2D fractured lattices at two different points in the system.
One was just before the lattices failed and a sample-spanning fracture was formed.
Figure 8.18 presents their results, indicating that each moment of the FD of frac-
tured lattices scales with the linear size L of the system with a different exponent.
The second point was at the maximum of the stress, where the system is entering
414 8. Brittle Fracture: The Discrete Approach
Figure 8.18. Rescaled moments m(q) = [M(q)/M(0)]1/2 versus the linear size L of the
lattice. The left figure gives the moments at the maximum current (where there is constant
gap scaling), whereas the right figure gives the moments right before the network fails (after
de Arcangelis et al., 1989).
the post-failure regime. In this case, the exponents that characterize the moments
of the FD near pc can be obtained from one of the exponents.
To understand better the difference between percolation and fractured lattices,
one important point to remember is that, while elastic properties of percolation
networks are controlled by the low moments of their FD (for example, the elastic
moduli are proportional to the second moment of the FD), fracture properties of
the same systems are controlled by the high moments of the FD. This is due to
the fact that fracture and breakdown occur where the largest loads (for example,
stress) are concentrated in the system, and the effect of such regions is manifested
only by the high moments of the FD.
Figure 8.19. Dependence of the ratio r = C11 /µ on the fraction of unbroken bonds during
fracture of a triangular lattice with stretching and bond-bending forces. The results, from
top to bottom, are for γ /α = 0.0, 0.01, 0.3 and 1 (after Sahimi and Arbabi 1992).
independently. In Figure 8.19 we present typical results for the ratio r = C11 /µ
as a function of the fraction of unbroken springs, for various values of γ /α, the
ratio of the elastic constants of the bond-bending model [see Eqs. (66) and (67)].
The last points of these curves represent C11 /µ right before the system failed
macroscopically. We refer to this as the incipient fracture point (IFP). As can be
seen, even though the initial states of the systems (i.e., their initial values of r
with no spring broken) are different, they all approach the same value of r as
the IFP is approached. Note also that, initially r remains essentially constant,
implying that the initial value of r is not sensitive to the presence of a few cracks
or even a collection of localized cracks. However, as damage accumulates and
the cracks grow, a turning point (TP) appears and r changes drastically. Because
γ /α = 0 corresponds to a lattice in which only central forces are present, Figure
8.19 indicates that this behavior is independent of the microscopic force law of the
system. The behavior of the system for γ /α = 1 is particularly interesting. Initially,
r remains essentially constant. However, as damage accumulates a TP appears
beyond which r decreases and reaches a minimum. But near the IFP, r rises again
416 8. Brittle Fracture: The Discrete Approach
and approaches its value at the IFP which appears to be universal. Simulations
(Sahimi and Arbabi, 1992) indicated that the value of r at IFP is universal and
independent of γ /α and ζ , the parameter of the threshold distribution, Eq. (70)
(unless, of course, ζ → 1). For 2D isotropic lattices (for example, a triangular
lattice) simulations indicated that
C11 5
. (102)
µ 4
The emergence of a universal fixed point at the IFP may mean that in a disordered
solid that undergoes quasi-static brittle fracture, the approach of r to its universal
value at the IFP may be interpreted as the “signature” of the material’s approach
to its global failure point. Although Figure 8.18 indicates that for certain values of
γ /α one may have a non-monotonic variation of r with the accumulated damage
(which, from an experimental view, implies that the closeness of r to its universal
value cannot be used for detecting the approach of the system to its global failure
point), for most real materials one has γ /α ≤ 0.3, and for such values of γ /α the
approach of r to its value at the IFP is always monotonic.
What is the theoretical explanation for the apparent universality of r? It is not
difficult to show that the dependence of C11 and µ on the fraction of unbroken
springs, as the IFP is approached, is similar to each other. As such, r represents an
amplitude ratio, and it is known (Aharony, 1980) from statistical mechanics that
certain amplitude ratios are universal. The apparent universality of r may mean
that, much like renormalization group theory of critical phenomena, universal
fixed points may be used for classifying various fracture processes. To do this,
we recall from Section 8.11.8 of Volume I that, it has been suggested (Bergman
and Kantor, 1984; Schwartz et al., 1985; Arbabi and Sahimi, 1988) that, in elastic
percolation networks C11 /µ takes on a universal value as pc is approached. For 2D
isotropic percolation networks near pc one has, C11 /µ 3, which is different from
the corresponding value for brittle fracture considered here. Sahimi and Goddard
(1986) suggested that brittle fracture of very heterogeneous networks is more or
less similar to a percolation process, and Roux et al. (1988) presented evidence
that in the limit of infinite disorder [i.e., the limit ζ → 1; see Eq. (70)] the quasi-
static lattice models of brittle fracture represent a type of percolation process.
For example, fracture in natural rock, a highly heterogeneous solid with scale-
dependent properties, may be a realization of this (Sahimi et al., 1993). On the
other hand, in most solids disorder is finite, and simulations indicated that even
for ζ = 0.8 the value of r at IFP is very different from that of elastic percolation
networks at pc , indicating that the limit ζ = 1 may be a type of singular point,
so that even for ζ = 1 − δ(δ 1), one should still obtain the value of r at the
IFP described here and not that of percolation networks at pc . Note that for 2D
isotropic systems, the Poisson’s ratio is given by νp = 1 − 2/r, implying that for
isotropic materials at the IFP, νp takes on a universal value.
Therefore, it has been proposed (Sahimi and Arbabi, 1992) that, value of the
Poisson’s point at the IFP may be used to classify various universality classes
8.2. Quasi-static Fracture of Heterogeneous Materials 417
as
log N
Df = (103)
log(1/)
where N is the number of fractures of length . Using Eq. (103) (the standard
box-counting method described in Chapter 1), fractal dimensions of the fractured
surfaces at Yucca Mountain were found to be in the range 1.5–1.7.
Another study was undertaken for the Geysers geothermal field in northeast
California (Sahimi et al., 1993). This field, from which heat and vapor are extracted
for use in power plants generating electrical power, covers an area of more than
35,000 acres and is one of the most significant geothermal fields in the world. Using
the box-counting method, Sahimi et al. (1993) determined the fractal dimension of
the fracture surfaces of the Geysers field and found, as did Barton and co-workers,
that at small length scales the fracture pattern is fractal with Df 1.5 − 1.7,
whereas at much larger length scales, Df 1.9. These results were interpreted
with the help of the quasi-static models of fracture described above (Sahimi et al.,
1993). In particular, note that this range of fractal dimension Df is essentially the
same as what one finds with the quasi-static lattice models of fracture.
There is also convincing experimental evidence indicating that in fractured rock
the Poisson’s ratio νp may take on a universal value at the IFP, in agreement with
the prediction of the quasi-static lattice models of brittle fracture described above.
The existence of a universal fixed point in fractured rock can be directly tested by
experimental measurements, since for 3D systems
$ %
Vp C11 1 1/2
= + , (104)
Vs µ 3
where Vp and Vs are the velocities of the shear and compressional waves in the
medium, respectively, which can be measured by established experimental proce-
dures (Brace and Orange, 1968). Sammonds et al. (1989) fractured four sandstone
samples at four different confining pressures, and measured Vp and Vs . Differ-
ent confining pressures result in different fracture patterns since they control the
closure of pre-existing cracks and nucleation and growth of new microcracks.
At the three lowest confining pressures the corresponding fracture patterns were
found to be brittle-like, and from their results one finds (Sahimi and Arbabi, 1992)
that Vp /Vs 1.14 ± 0.04 at the IFP for all the three fractured sandstones, imply-
ing a universal value for C11 /µ. At the highest confining pressure fracture was
ductile-like, and although, as expected, the stress-strain diagrams of the sample
was not similar to that of brittle fracture, their results indicated that even for this
case Vp /Vs 1.1, beyond the point at which stress became independent of strain
(which is typical of stress-strain diagrams of ductile fracture), consistent again
with the value for brittle fracture at the lower confining pressures. These data pro-
vide strong experimental support for the existence of universal fixed points at the
IFP. Note that since for 3D isotropic systems, we have
3(C11 /µ − 1)
νp = , (105)
2 + 6(C11 /µ − 1)
8.3. Dynamic Brittle Fracture 419
the experimental data of Sammonds et al. (1989) imply that for their samples, νp
0.1 at the IFP. On the other hand, if we use C11 /µ 4/3, the corresponding value
for 3D isotropic elastic percolation networks at pc , we find νp 1/4. Thus, over
the length scale used in these experiments, the sandstones examined by Sammonds
et al. (1989) must have been relatively homogeneous, since their Poisson’s ratio at
the IFP (νp 0.1) is different from that of percolation networks at pc (νp 1/4),
which corresponds to highly heterogeneous systems. These results also support the
validity of classifying fracture processes according to the value of the Poisson’s
ratio at the IFP.
One may wonder about the appropriateness of such models for the dynamic
behavior of a crack in a crystal. The critical question, which could have also been
asked about the lattice models of quasi-static fracture is, is the lattice essential,
or can one make the underlying lattice go away by taking a continuum limit? To
our knowledge, all attempts for describing the cohesive zone of brittle materials
in a continuum framework have run into severe difficulties (see, for example,
the discussion by Langer and Lobkovsky, 1998). These problems do not arise in
lattice models described here, nor are they encountered in the atomic-scale MD
simulations that will be described in Chapter 9.
The simplicity of the ideal brittle crystal is somewhat misleading in a number
of respects. Therefore, before introducing the models, a few natural questions
regarding the generality of their predictions are posed and commented on.
(1) Does the simple force law employed between the nodes of the lattice (the
atoms in the crystal) neglect any essential aspect of the dynamics? Experience
with the MD simulations of dynamic fracture, to be described in Chapter 9,
indicates that the same qualitative results are also observed in a brittle ideal
crystal (that is, one without disorder). Moreover, we already know that, the
same type of force law, when utilized in the quasi-static case, produced very
insightful results.
(2) The most interesting calculations that have been carried out so far (see be-
low) are for a strip geometry. Is this geometry too restrictive? The general
formulation of fracture mechanics provides an answer to this question, since
it tells us that as long as the conditions of small-scale yielding are satisfied,
the behavior of a crack is entirely governed by the structure of the stress fields
in the near vicinity of the crack tip. These fields are solely determined by the
flux of energy to the crack tip. A given energy flux can be provided by an
infinite number of different loading configurations, but the resulting dynamics
of the crack will all be the same. As a result, the specific geometry used to
load the system is irrelevant and no generality is lost by the use of a specific
loading configuration. This fact is born out not only by the experimental work
described in Chapters 6 and 7, but also by the lattice models of quasi-static
fracture described earlier in this chapter.
(3) Are the predictions for a perfect (without disorder) crystal or lattice relevant
to amorphous materials? This is still an open question. However, the results of
the lattice calculations appear to be remarkably robust. Adding weak quenched
disorder to the crystal has little qualitative effect (recall that the same conclu-
sion was reached with the lattice models of quasi-static fracture). The effects
of topological disorder are not known. However, it is well-known (see, for
example, Holland and Marder, 1998) that when the temperature of a brittle
crystal increases above zero, the velocity gap (see below) becomes narrower,
and its behavior is reminiscent of what is seen in experiments performed on
amorphous materials, which were described in Chapters 6 and 7.
One of the earliest exact calculation with lattice models of dynamic fracture
was carried out by Kulakhmetova et al. (1984) who showed that it is possible to
8.3. Dynamic Brittle Fracture 421
find exact analytical expressions for Mode I fractures moving in a square lattice.
They derived exact relationships between the energy flux to a crack tip and its
velocity, observed that phonons must be emitted by moving cracks, and calcu-
lated their frequencies and amplitudes. Later calculations by Marder and Gross
(1995) extended these results to other lattice geometries, allowed for a general
Poisson’s ratio, showed that there is a minimum allowed crack velocity (which
was found when steady fracture motion was linearly stable), computed the point
at which steady motion becomes unstable to a branching instability, and estimated
the spacing between branches. These calculations are extremely elaborate, with
the analytical expressions being so lengthy that they do not even fit on printed
pages (see Gross, 1995)! This is particularly true about the calculations for Mode
I equations. For this reason, we summarize some of the results for Mode I, and
then proceed to describe in detail how the calculations are carried out in the case
of anti-plane shear, Mode III, where the algebra is much less demanding, but most
of the ideas are the same. But, let us first describe a relatively simple lattice model
of dynamic fracture which could predict some aspects of dynamic fracture.
One of the earliest numerical simulation of dynamic fracture, based on a lattice
model, was carried out by Mori et al. (1991). A triangular lattice was used in their
work, each bond of which was assumed to be a Hooke’s spring, if the threshold
for its breaking was not exceeded. They assumed that a spring breaks irreversibly
if it is stretched beyond a given threshold. Each spring was also characterized by
a spring constant α. The nodes of the lattice were occupied by particles of mass
m. Suppose that e is the elongation vector of the springs that are connected to a
particle at position Ri . The equation of motion for the particle at time t is
d 2 Ri dRi
m = −D − ke, (106)
dt 2 dt
where the first term on the right-hand side is a damping term with D being the
damping constant; this term essentially represents some type of friction. D is not
a parameter that can be measured easily in any experiment, and thus should be
treated as a free parameter of the model. Setting dRi /dt = vi (where vi is the
velocity of node i), we obtain two equations that govern Ri and vi which, when
written in a finite-difference form, are given by
$ %
D
vi (t + t) = 1 − t vi (t) − t ke, (107)
m
t
Ri (t + t) = Ri (t) + vi (t), (108)
m
where t is the time increment.
To begin the simulations, an initial microcrack is inserted in the middle of the
lattice. As the initial condition, the lattice is stretched by an amount . Equations
(107) and (108) are then solved at time t and the springs are examined to identify
those that have exceeded their threshold. Such springs are broken, the time t is
advanced by t, Eqs. (107) and (108) are solved again, the next springs(s) is (are)
broken, and so on. An important parameter of the model is I = 0 /, where 0
422 8. Brittle Fracture: The Discrete Approach
Figure 8.20. Fracture pattern in the dynamic model of Mori et al. (1991) for I = 1.0083.
The arrow indicates the location of the initial microcrack.
is the initial length of the springs, and is the initial amount of stretching that
the lattice has suffered. Increasing I is, in some sense, equivalent to increasing
the temperature of the system. Complex fracture pattern can emerge, depending
on the value of I . It was found that for I < 1.0085 the fracture pattern was tree-
like, and in fact no microcrack was formed if no initial crack was inserted in the
system. However, for I > 1.0085 the microcracks became connected and formed
a network and, moreover, even with no initial microcrack in the system, fractures
were formed “spontaneously.” Figure 8.20 shows the typical fracture pattern for
I = 1.0083.
In what follows, we describe the exact calculations carried out by Marder and
co-worker. Some of our discussions follow closely that of Fineberg and Marder
(1999), while the rest are based on the review by Sahimi (1998).
into a component along d̂ 1 , which is along the line that connects two neighboring
atoms, and a component that is along d̂⊥1 which is perpendicular to it. Hence, the
first component corresponds to central forces between atoms, whereas the second
component is a non-central force. Recall from Chapter 8 of Volume I (see also
Chapter 9) that in real materials non-central forces between atoms are the rule
rather than the exception. Suppose now that the restoring force parallel to the
direction of equilibrium bonds is proportional to F , while that perpendicular to
this direction is proportional to F⊥ . Therefore, if U1 = (U1x , U1y ), then the force
due to the displacement of the particle along U1 = ui−1,j +1 − ui,j is given by
√ √
−1 3 −1 3
F d̂ 1 (U1 · d̂ 1 ) + F⊥ d̂⊥1 (U1 · d̂⊥1 ) = F⊥ , U1x , U1y
2 2 2 2
√ √
3 1 3 1
+F⊥ , U1x , U1y . (109)
2 2 2 2
If, by a similar method, one adds up contributions from other particles, one obtains
for the force due to neighbors
6
F(m, n) = Fq d̂qj [Uj (m, n) · d̂qj ] (110)
j =1 q= ,⊥
By varying the constants F and F⊥ , one can obtain any desired values of shear
and longitudinal wave speeds, which are given by
3a 2
cl2 = (F⊥ + 3F ), (111)
8m
3a 2
ct2 = (3F⊥ + F ), (112)
8m
where m is the mass of each particle. In addition to the forces between the neigh-
boring atoms, it is possible to take into account the effect of complex dissipative
functions that depend upon particle velocities. Marder and Gross (1995) added a
term to the equations so as to reproduce the experimentally-measured frequency
dependence of sound attenuation in Plexiglas. There is a slight technical restriction
in the calculations of Marder and Gross (1995) in that, forces right on the crack
line are required to be central. However, more detailed calculations indicate that,
when this technical limitation is removed, the results do not change significantly.
Figure 8.21 shows schematics of steady-state propagation of a crack in the lattice
loaded in Mode I.
Detailed, lengthy and difficult calculations show that, in Mode I loading, many
details of the relation between loading and fracture velocity depend upon ratios of
the sound speeds and also the frequency dependence of dissipation, implying that
there exists no universal curve that can describe Mode I fracture. When only central
forces are present (F⊥ = 0), it is difficult (but not impossible) to have steady-state
fracture propagation, implying that the range of loads for which cracks can propa-
424 8. Brittle Fracture: The Discrete Approach
gate in a stable fashion is small, and depends upon the amount of dissipation. Thus,
the presence of non-central forces is essential to having stable fracture growth. This
also has important implications for MD simulations of dynamic fracture propa-
gation in materials. For example, the proper form of the interatomic potential for
representing the non-central forces, such as the bond-bending ones, is an important
issue in the MD simulations which will be discussed in detail in Chapter 9.
Given the complexities of exact calculations for Mode I fracture, it is wiser to
consider a simpler geometry for which the analytical methods can be discussed in
complete detail, and all the essential ideas that are needed for understanding Mode
I can be explored with much less elaborate calculations. This is the subject of the
next section.
Figure 8.22. Dynamic fracture of triangular crystal in anti-plane shear (after Fineberg and
Marder, 1999).
difference in the height between them. Hence, in terms of u(m, n), the equation
of motion of the system is given by
d 2u du 1
2
= −b + F[u(m , n ) − u(m, n)], (113)
dt dt 2
m ,n
describes ideal brittle springs, with # being the step function. The boundary
condition is given by
1
u[m, ±(N + )] = ±uN . (115)
2
It is important to determine the value of uN for which there is just enough stored
energy per unit length to the right of the fracture to break the pair of bonds connected
to each lattice site on the crack line. For m 0 one has
nuN
u(m, n) = , (116)
N+ 1
2
and the height difference between mass points with adjacent values of n is
nN
Uright = . (117)
N+ 1
2
426 8. Brittle Fracture: The Discrete Approach
Therefore, the energy stored per unit length in the (2N + 1) rows of bonds is
1
× (2 upper bonds/site) × (rows) × (spring constant) × Uright
2
2 2
1 1 uN
= × 2(2N + l) × × = 2N0 u2N , (118)
2 2 N+2 1
where N0 = 1/(2N + 1). The energy required to break two bonds each time the
crack propagates by a unit length is given by
1
× (2 bonds/site) × (spring constant) × (separation at fracture)2
2
1 1
= × 2 × × (2uf )2 = 2u2f . (119)
2 2
Therefore, by equating Eqs. (118) and (119) one obtains the proper dimensionless
measure of the external driving force,
√
uN N0
= , (120)
uf
which reaches 1 as soon as there is enough energy to the right of the crack to break
the bonds along the crack line. Note that is linearly related to the displacement
uN imposed at the edges of the strip.
Slepyan (1981) and Kulakhmetova et al. (1984) developed the techniques for
solving problems of this type. However, there are some differences between details
of their solution and what is presented here, because Eq. (113) describes motion
in a strip, which is what is considered here, rather than an infinite plate that they
studied. Both geometries have certain advantages. The strip is preferable to the
infinite plate if one wishes to compare the predictions with the results of numerical
simulations. On the other hand, using the infinite plate results in certain natural
limits. Moreover, the lattice considered here is a triangular rather than a square
lattice used by Slepyan and co-workers.
We assume that a fracture moves in steady state, so that one by one, the bonds
connecting u(m, 12 ) with u(m + 1, − 12 ) or u(m, − 12 ) break because, as a result of
the driving force described by Eq. (115), the distance between these atoms exceeds
the limit set by Eq. (115). Assuming that the times at which the bonds break are
known, the original nonlinear problem is immediately transformed into a linear
problem. However, once the solution of the linear problem has been derived, one
must verify that,
(1) bonds break at the time that they are supposed to. If we impose this condition,
we obtain a relation between the crack velocity v and dimensionless loading .
(2) Conversely, no bonds break when they are not supposed to. This condition
results in a velocity gap on the low end of the velocity range, and leads to
crack tip instabilities above a critical energy flux (see below).
We note here that steady states in a perfect lattice or crystal are more complex
than those in a continuum, since in the latter case a steady state acts as u(x, vt),
8.3. Dynamic Brittle Fracture 427
whereas the closest to such a state that one can come in a triangular lattice is by
having the symmetry
u(m, n, t) = u(m + 1, n, t + 1/v), (121)
and also
$ %
1
u(m, n, t) = −u m, −n, t − − gn /v , (122)
2
which implies in particular that
1 1
u(m, , t) = −u m, − , t − 1/(2v) . (123)
2 2
Here
⎧
⎪
⎪ 0 if n = 12 , 52 · · ·
⎨
gn = 1 if n = 32 , 72 · · · (124)
⎪
⎪
⎩
mod(n − 12 , 2) in general.
By assuming that a fracture is in steady state, one can eliminate the variable m
entirely from the equation of motion. We define
un (t) = u(0, n, t), (125)
and write the equations of motion in steady state as
⎡ ⎤
2
un+1 [t − (gn+1 − 1)/v] +un+1 (t − gn+1 /v)
d un 1⎢ ⎥ dun
= ⎣+un (t + 1/v) −6un (t) + un (t − 1/v)⎦ − b
dt 2 2 dt
+un−1 [t − (gn−1 − 1)/v] +un−1 (t − gn−1 /v)
(126)
if n > 1/2. For n = 1/2 we have
d 2 u1/2
=
dt 2
⎡ ⎤
u3/2 (t) +u3/2 (t − 1/v)
1⎢ ⎢+u1/2 (t + 1/v) − 4u1/2 (t)
⎥
⎥ − b du1/2
+u1/2 (t − 1/v)
2⎣ ⎦ dt
+[u−1/2 (t) − u1/2 (t)]#(−t) +[u−1/2 (t − 1/v) − u1/2 (t)]#[1/(2v) − t]
(127)
We have assumed that t = 0 is the time at which the bond between u(0, 12 , t)
and u(0, − 12 , t) breaks, and therefore, by symmetry, the time at which the bond
between u(0, 12 , t) and u(1, − 12 , t) breaks is 1/(2v).
Above the fracture line, Eqs. (126) and (127) are completely linear, and thus
it is not difficult to derive the solution that yields the motion of every atom with
n > 1/2 in terms of the behavior of an atom with n = 1/2. If we Fourier transform
428 8. Brittle Fracture: The Discrete Approach
(134)
This solution equals u1/2 for n = 1/2, and equals 2αuN /(α 2 + ω2 ) for n = N + 12 .
Introducing α is necessitated by the fact that for n = N + 12 , u(m, n, t) = uN .
Instead of working with the Fourier transform of this boundary condition, which
is a delta function and difficult to work with, it is better to use the following
boundary condition
uN +1/2 (t) = uN e−α|t| , (135)
and let α → 0 at the end of the calculation. In the analysis that follows, frequent
use is made of the fact that α is small.
8.3. Dynamic Brittle Fracture 429
From a physical point of view, the most interesting variable is not u1/2 , but the
distance between the bonds which will actually break. Hence, we define
u1/2 (t) − u−1/2 (t) u1/2 (t) + u1/2 (t + 1/2v)
U (t) = = , (136)
2 2
to rewrite Eq. (127) as
⎡ ⎤
u3/2 (t) +u3/2 (t − 1/v)
2 1⎢
d u1/2 ⎥ du1/2
= ⎣+u1/2 (t + 1/v) − 4 +u1/2 (t) + u1/2 (t − 1/v) ⎦−b .
dt 2 2 dt
−2U (t)#(−t) −2U (t − 1/2v)#[1/(2v) − t]
(137)
If we Fourier transform this expression, use Eq. (134), and define
∞
U ± (ω) = U (t)#(±t) exp(iωt)dω, (138)
−∞
we obtain
uN 2α
u1/2 (ω)F (ω) − (1 + eiω/2v )U − (ω) = − , (139)
N ω + α2
2
where
y [N −1] − y −[N −1]
F (ω) = − 2z cos(ω/2v) + 1. (140)
y N − y −N
We now use Eq. (136) in the form
1
U (ω) = (1 + e−iω/2v )u1/2 (ω) (141)
2
to obtain
uN 2α
U (ω)F (ω) − 2 cos2 (ω/4v)U − (ω) = − . (142)
N ω + α2
2
If we write
U (ω) = U + (ω) + U − (ω) (143)
we finally obtain
$ %
1 1
U + (ω)Q(ω) + U − (ω) = uN N0 + , (144)
α + iω α − iω
where
F (ω)
Q(ω) = . (145)
F (ω) − 2 cos2 (ω/4v)
To derive the right-hand side of Eq. (144), we used the facts that F (0) = −1/N ,
and that α is very small, so that the right-hand side of this equation is a delta
function.
430 8. Brittle Fracture: The Discrete Approach
We now utilize the Wiener–Hopf technique (see, for example, Noble, 1958) to
write
Q− (ω)
Q(ω) = , (146)
Q+ (ω)
where Q− and Q+ are free of poles and zeroes in the lower and upper complex
ω planes, respectively. This decomposition can be carried out with the explicit
formula
1 ln Q(ω )
Q± (ω) = exp lim dω
. (147)
ε→0 2π iω ∓ ε − iω
We now separate Eq. (144) into two parts, one of which has poles only in the lower
half plane, while the other part has poles only in the upper half plane:
U + (ω) u N N0 1 u N N0 1 U − (ω)
− = − . (148)
Q+ (ω) Q− (0) −iω + α Q− (0) iω + α Q− (ω)
Because the right- and left-hand sides of Eq. (148) have poles in opposite sections
of the complex plane, they must separately equal a constant. However, the constant
must be zero, because otherwise U − and U + will behave as a delta function near
t = 0, and therefore
uN N0 Q− (ω)
U − (ω) = , (149)
Q− (0)(α + iω)
and
uN N0 Q+ (ω)
U + (ω) = , (150)
Q− (0)(α − iω)
which provide an explicit solution for U (ω). Numerical evaluation of Eq. (147),
and U (t) using Eqs. (149) and (150) is fairly straightforward (using, for example,
fast Fourier transforms). However, in carrying out the numerical transforms, one
must carefully analyze the behavior of the functions for large values of ω. If these
functions decay as 1/(iω) [the inverse Fourier transform of which is #(t)], we
should subtract the 1/(iω) off before the numerical transform is performed, after
which we should add this term back with the appropriate step function. Conversely,
if the functions to be Fourier transformed have #(t) discontinuity, it is best to
subtract off the appropriate multiple of et #(t) before Fourier transforming, and
then add on the appropriate multiple of 1/(1 − iω). A solution of Eqs. (149) and
(150) constructed in this manner is given in Figure 8.23.
One can now derive a relation between the dimensionless displacement and
the crack velocity v. Recall that making the transition from the original nonlinear
problem posed by Eq. (113) to the linear problem expressed by Eq. (126) relies on
assuming that bonds along the crack line break at time intervals of 1/2v. Because
of the symmetries expressed by Eqs. (121)–(123), it is sufficient to guarantee that
u(t = 0) = uf . (151)
8.3. Dynamic Brittle Fracture 431
Figure 8.23. Predictions of Eqs. (149) and (150) for v = 0.5, N = 9 and b = 0.01 (after
Fineberg and Marder, 1999).
All the displacements are proportional to the boundary displacement uN , and hence
Eq. (151) fixes a unique value of uN and its dimensionless counterpart, . This
means that, once one assumes that the fracture moves in steady state at a velocity
v, there is a unique to make it possible.
To derive Eq. (151), we must require that
1
lim U − (ω) exp(−iωt)dω = uf . (152)
t→0− 2π
and that any function that for large ω behaves as 1/(iω) has a step function
discontinuity at the origin. Therefore, Eqs. (149) and (152) become
Q− (∞)
uf = uN N0 . (154)
Q− (0)
Since Eq. (145) yields Q(∞) = 1, one obtains from Eq. (147) that,
Q− (0)
= √ . (156)
N0
432 8. Brittle Fracture: The Discrete Approach
To make this result more explicit, we use Eq. (147) and the fact that
1 1 ln Q(ω ) ln Q(−ω )
Q− (0) = exp + dω
2π 2 ε − iω ε + iω
(157)
1 1 Q(ω ) ε
= exp ln + ln Q(0) dω .
2π −2iω Q̄(ω ) ε 2 + ω2
Therefore,
!
1 1 Q(ω )
Q− (0) = N0 exp − ln dω . (158)
2π 2iω Q(ω )
Inserting Eq. (158) into Eq. (156) yields
1 1 Q(ω )
= exp − ln dω . (159)
2π 2iω Q(ω )
In order to obtain an expression that is correct not only for Mode III model
considered here, but also for more general cases, we rewrite Eq. (159) as
1 1 )}dω ,
= C exp − {ln Q(ω ) − ln Q(ω (160)
2π 2iω
where C is a constant of order unity that is determined by the geometry of the
√ example, for the triangular lattice loaded in Mode III, C = 1, whereas
lattice. For
C = 2/ 3 for the same lattice loaded in Mode I (Marder and Gross, 1995). The
advantage of Eq. (160) is that it is suitable for numerical evaluation, since there is
no uncertainty regarding the phase of the logarithm.
When b, the coefficient of the dissipation term, becomes sufficiently small, Q
is real for real ω except in the small neighborhood of the isolated roots and poles
that are near the real ω-axis. Suppose that ri+ are the roots of Q with negative
imaginary part (since they belong to Q+ ), ri− are the roots of Q with positive
imaginary part, and pi± are the poles of Q. Equation (160) can be written as
#'
=C ri− pi+ /ri+ pi− , for b → 0. (161)
i
In deriving Eq. (161) use was made of the fact that, away from a root or pole of
Q, the integrand of Eq. (160) vanishes. Together with Eqs. (149) and Eqs. (150),
Eqs. (160) and (161) constitute the formal solution of the model. Since Q is a func-
tion of the steady-state velocity v, Eq. (159) relates the external driving force im-
posed on the crystal, , to the velocity v of the crack. The results of a typical calcu-
lation for N = 9 are presented in Figure 8.24, which show clearly the velocity gap.
Having determined the formal solution of the model, we can now investigate a
few important issues.
√
Figure 8.24. Fracture velocity v, normalized by sound velocity vs = 3/2, versus the
driving force , using N = 9. The end points of the lower curve indicate the linearly-stable
lattice-trapped states (after Fineberg and Marder, 1999).
all the energy stored to the right of the fracture tip will be used for breaking bonds.
What happens to the remaining energy depends on the amount of dissipation b,
and the distance from the fracture tip at which one inspects the system. In the
limit b → 0, travelling waves leave the fracture tip and carry energy off in its
wake, with the amount of energy that they carry off being independent of b. For all
non-zero values of b, however, the travelling waves will eventually decay, and the
extra energy will have been absorbed by dissipation, but the value of b determines
whether one views the process as microscopic or macroscopic.
The frequencies of the radiation emitted by the crack have a simple physical
interpretation. Consider the motion of a particle through a lattice, in which the
phonons are described by the dispersion relation, ωα (k). If the particle moves
with a constant velocity v and interacts with the various ions according to some
function F, then, to linear order, the motions of the ions can be described by a
matrix M which describes their interactions with each other as
(l)
d 2 ui
(n) (n)
m = − M ij (R (l)
− R )u j + Fi (R (n) − vt). (162)
dt 2 n
j,n
(l)
Multiplying both sides of Eq. (162) by eik·R , summing over l, and letting K and
be, respectively, the reciprocal lattice vectors and the volume of a unit cell, yield
d 2 ui (k) 1
m 2
= Mij (k)uj (k) + exp[iv · (k + K)t]Fi (k + K). (163)
dt
j K
434 8. Brittle Fracture: The Discrete Approach
By inspection we can show that the lattice frequencies excited in this way are those
which in the extended zone scheme satisfy the following equation (Ashcroft and
Mermin, 1976)
ω(k) = v · k. (164)
If we pretend that the crack is a particle, we can use Eq. (164) to predict the phonons
that the crack emits.
There are actually two phonon dispersion relations to consider, one for propa-
gating radiation far behind the crack tip, and the other for propagating radiation far
ahead of the tip. Far behind the crack tip, all the bonds are broken, and therefore
to find the travelling one we must set U − = 0 in Eq. (142), since all the bonds
behind the tip are broken, and also uN = 0, because phonons can propagate with-
out any driving term, which then lead us to F (ω) = 0. Similarly, because no bond
far ahead of the crack tip is broken, U − = U , and the condition for phonons is
F (ω) − 2 cos2 (ω/4v) = 0. Therefore, according to Eq. (145), the roots and poles
of Q(ω) are the phonon frequencies behind and ahead of the fracture, which are
also the quantities that appear in Eq. (161).
can be tested by using the numerical solutions of Eqs. (149) and (150), since the
solution fails above a critical
√ value c of . The sound speed vs (= cR ) equals,
in dimensionless units, 3/2, and thus we rescale velocities by this value. For
example, for N = 9, at a velocity, vc /vs = 0.666 · · ·, c = 1.158 · · ·, and the
bond between u(0, 12 ) and u(1, 12 ) reaches a distance of 2uf shortly after the bond
between u(0, 12 ) and u(0, − 12 ) breaks. The steady-state solutions that are obtained
when the lattice is strained with larger values of are inconsistent; only dynamical
solutions more complex than steady states, involving the breaking of bonds off
the fracture path, are possible. To investigate these states, one must numerically
solve Eq. (113). Such numerical simulations have been carried out (Marder and
Liu, 1993). The results show that just above the threshold at which horizontal
bonds begin to break, the distance between these extra broken bonds diverges.
The reason is that breaking a horizontal bond takes energy from the fracture and
slows it down below the critical value. The fracture then tries once more to reach
the steady state, and only in the last stages of the approach does another horizontal
bond break, hence beginning the process again.
A rough estimate of the distance between broken horizontal bonds can be ob-
tained as follows. Let h (t) be the length of an endangered horizontal bond. One
must view the problem in a reference frame moving with the fracture tip, and there-
fore at every time interval 1/v one shifts attention to a bond one lattice spacing
to the right. When is only slightly greater than c , the length of such a bond,
viewed in a moving frame, should behave, before it breaks, as
∂h
h ∼ 2uf + ( − c ) − δ exp(−bt). (165)
∂
Here, ∂h /∂ denotes the rate at which the steady-state length of h would change
with , if this bond were not allowed to break, and δ describes how much smaller
than its steady-state value the bond ends up after the breaking event occurs. Marder
and Gross (1995) showed that deviations from steady states die away at long times
as exp(−bt). From Eq. (165) we can estimate the time between breaking events
by setting h = 2uf and solving the equation for t. The result is that the frequency
f with which horizontal bonds break should scale above the critical strain c as
b
f ∼− . (166)
ln[(δ)−1 ( − c )∂h /∂]
However, the accuracy of Eq. (166) has proven to be difficult to check.
(1) The force law between the atoms is actually much more complex than ideal
breaking bonds. Gao (1993) has pointed out that the Rayleigh wave speed cR
in the vicinity of a crack tip may be significantly lower than its value far away
from the tip, because material is being stretched beyond the range of validity
of linear elasticity.
(2) The experiments are at room temperature, while the calculations are at zero
temperature.
(3) The experiments are in amorphous materials, while the calculations are for a
crystal or perfect lattice.
Before closing this section, let us summarize the main results obtained from the
ideal brittle crystal.
(i) For a range of loads above the Griffith point, a fracture can be trapped by
the crystal, i.e., it neither moves nor heals, although it is energetically possible for
the fracture to move (Thomson et al., 1971; Thomson, 1986). However, molecular
dynamics simulations, to be described in Chapter 9, indicate that lattice trapping
occurs at very low temperatures and in fact disappears at room temperature.
(ii) Steady-state fracture motion exists, and is a stable attractor for a range of
energy flux.
(iii) Steadily-moving fractures emit phonons with frequencies that can be
computed from a simple conservation law.
(iv) The relation between the fracture energy and velocity can be computed.
(v) The slowest steady state runs at around 0.20cR . There is no slower-moving
steady-state fracture, and therefore there is a velocity gap.
(vi) At an upper critical energy flux, steady-state fractures become unstable and
generate frustrated branching events in a fashion reminiscent of experiments in
amorphous materials described in Chapter 7.
Furukawa (1993) modified the model proposed by Mori et al. (1991) (see Section
8.3) by including a shear friction in the model. Thus, in this model the equation of
motion is given by
dvi
m = −D1 vi − D2 (vi − vj ) + F(Rj − Ri ), (167)
dt
j j
where the second term on the right-hand side represents shear friction or dissipa-
tion, while the last term is the force term which was taken to be F(R) = Rf (R)/R,
where R is the magnitude of R. The function f (R) was selected to be f (R) =
R − 1 for R ≤ R0 , and f (R) = (R0 − 1) exp[−κ(R − R0 )] for R ≥ R0 , where
R0 and κ are two constants. A bond breaks if R > R0 . Both square and triangular
lattices were used. Fracture formation was initiated by inserting an initial microc-
rack at the center of the lattice. In one case the lattice spacing in the direction of the
macroscopic deformation was 1, and the remaining lattice distances were Re with
1 < Re < R0 . An interesting prediction of the model was that, when κ = ∞, the
fracture velocity v = /t, where is the distance between the central microcrack
and the most distant broken bond at time t, follows a power law:
(Re − 1)x1
v∼ , (168)
(R0 − Re )x2
where the velocity has been scaled by the 1D sound velocity. In most cases
x1 = x2 = 1, except in the square lattice without the dissipation term (D2 = 0),
in which case x1 = 1 and x2 = 1/3. In this case fracture propagation was sub-
sonic, whereas it was supersonic on the triangular lattice if (R0 − Re )/(Re − 1)
was small (see Sections 7.7.1, 7.8.15, and 9.8.3.3 for discussions of supersonic
fracture propagation). In the second case that was studied, all the lattice spacings
were equal. A variety of interesting fracture patterns were obtained, some of which
are shown in Figure 8.25. Also obtained were oscillatory fracture patterns, some
of which are also shown in Figure 8.25. As discussed above and in Chapters 6 and
7, such oscillatory fracture patterns contribute to the dynamic instability observed
during fracture propagation.
Another interesting deterministic lattice model of dynamic fracture was pro-
posed by Rautiainen et al. (1995). In their model each bond ij of a square lattice
was an elastic element with an interaction energy that was described by
α γ
Hij = [(ui − uj ) · R ]2 + [(ui − uj ) · R⊥ ]2 , (169)
2 2
where ui is the displacement of node i, and R and R⊥ are the unit vectors paral-
lel and perpendicular to the vector connecting i and j in the undeformed lattice,
respectively. To include disorder in the model, a fraction q of the bonds was re-
moved at random before fracture simulations began. To introduce dynamics into
the model, a dissipation mechanism was included in the model by incorporating in
it Maxwellian viscoelasticity which allows the description of relaxation and dis-
sipation of elastic energy as a dynamical decay of the local forces. The constitutive
438 8. Brittle Fracture: The Discrete Approach
! "
# $ % &
Figure 8.25. Regular and oscillatory fracture patterns obtained by the dynamic model of
Furukawa (1993). The regular patterns were obtained under isotropic tension.
equation for the forces acting at each bond at time t was taken to be
dFij dFH 1
= − Fij , (170)
dt dt tr
where Fij is the force between i and j arising out of their interaction, FH is
the elastic force derived from Eq. (169), and tr is a phenomenological parameter
which can be considered as a relaxation time scale. In effect, each bond is replaced
by a Maxwellian viscoelastic element—a spring and a dashpot in series, and was
considered as broken if its length exceeded a critical value.
The model predicts brittle fracture in the limit of slow straining. At finite strain
rates, damage development becomes ductile with increasing dissipation. For small
tr and q, the number of broken bonds increases rapidly at the initial stages of the
fracture history. However, after some time damage accumulation stops and the sys-
tem resists rupture. This is due to local viscoelastic dissipation which arrests crack
growth. For large tr the number of broken bonds increases slowly, and the manner
by which the bonds break is correlated. Thus, ductility increases with decreasing
tr . Hassold and Srolovitz (1989) had already shown, using a Born and the quasi-
static lattice model described in Section 8.2, that crack arrest can be controlled
by varying α/γ . Therefore, the role of α/γ in the quasi-static model is played by
the relaxation parameter tr . The crack velocity was found to be approximately,
8.3. Dynamic Brittle Fracture 439
Figure 8.26. Fracture patterns in the dynamic model of Heino and Kaski (1996). The top
pattern corresponds to tr = 0, while the rest are for tr = 0.
√
v γ . A dynamic Born model was developed by Mart́in et al. (2000) that also
produced many of these predictions.
In another lattice model of dynamic fracture, Heino and Kaski (1996) used the
same Hamiltonian as in Eq. (169), but the lattice was more complex. Each bond of
the lattice consisted of two perpendicular springs in fixed directions representing
tensile and bending behavior. A dashpot was connected in series to each spring. In
a tensile experiment, the sites move which results in elongation or shrinking of the
springs and dashpots. Disorder was introduced into the model by assuming that
each bond has its own α and distributing it according to a uniform distribution. The
ratio α/γ for each bond was held constant though, which corresponds to varying
the Young’s modulus of the bonds. A bond was considered as broken when its
strain exceeded a threshold. Interesting fracture patterns were generated by the
model by varying its parameters, some of which are shown in Figure 8.26. As can
be seen, there is a dominating fracture that eventually spans the system and causes
it to fail. However, many daughter cracks also appear essentially symmetrically
on both sides of the main fracture. They appear periodically and advance a short
distance before dissipation damps their growth. If the network is more disordered,
the periodicity disappears and the fracture pattern becomes irregular.
However, the most interesting result emerging from this model was the behavior
of the crack tip velocity as a function of the crack width. Initially, the velocity
increased rapidly, corresponding to the emergence of a straight fracture. However,
after some time the oscillatory daughter fractures appeared, and thus the velocity
also started to oscillate with the crack width, and hence with the time. Increasing
α increased the crack speed and its oscillation frequency, but decreased the length
of the daughter fractures, although the angle that they made with the main fracture
was unaffected by α. Heino and Kaski (1997) combined a finite-element method
(for discretizing the continuum elasticity equations and obtaining a finite-element
440 8. Brittle Fracture: The Discrete Approach
mesh) with the bond-breaking process used in lattice models. The model provided
the same type of predictions as those of their earlier model.
Åström and Timonen (1996) used a square lattice of beams (see Section 8.2)
to study dynamic fracture. Their model was the dynamic analogue of the lattice
models of quasi-static brittle fracture described in 8.2. The network was strained
by an amount in the y direction. The sites at the top and bottom edges of the
network were constrained to remain at their original positions, while the sites on
the left and right edges were free to move without constraints. The dynamics of the
model was calculated using a discrete form of the Newton’s equation of motion
including a linear viscous dissipation term (see also Section 8.4 below). A beam
was considered as broken if the strain on it exceeded a pre-assigned threshold
value. Åström and Timonen (1996) showed that, by tuning the strain and the ratio
of axial to bending stiffness of the beams, a fracture can propagate either straight,
or branch, or bifurcate. The fact that such features can be obtained by both the
Born model of Heino and Kaski (1996), described above, and by the beam model
of Åström and Timonen, indicates the universality of such features. Moreover, for
the branching fracture, Åström and Timonen (1996) found that their trajectories
follow a power law,
y ∼ x 0.7 , (171)
where x and y are, respectively, the directions parallel and perpendicular to the
direction of the main crack, with the origin being the point at which the micro-
branch begins. This result is in complete agreement with the power-law observed
in experiments in both polymethylmethacrylate (PMMA) and glass (see Section
7.10.4).
pattern is formed. For v > vc the amplitude of the oscillations depends linearly
on the mean velocity of the propagating fracture. Thus, the motion of fractures is
governed by a dynamical instability, and explains why the velocity of their tip does
not attain the limiting Rayleigh velocity cR predicted by the linear elasticity theory.
In another set of beautiful experiments, Gross et al. (1993) used two materials,
the PMMA and soda-lime glass, to show that all features of dynamics of crack
propagation in the two materials, such as acoustic emission, crack velocity, and
surface structure, exhibit quantitative similarity with each other. Thus, there exist
universal characteristics of fracture energy in most materials that are the result of
energy dissipation in a dynamical instability. Perhaps the most spectacular exper-
iments were carried out by Sharon et al. (1995) and Sharon and Fineberg (1996)
using the brittle plastic PMMA. They identified the origin of the dynamical in-
stability during fracture propagation as being the nucleation and growth of the
daughter cracks which limit the speed of the propagating crack tip. The daughter
fracture carries away a fraction of the energy concentrated at the tip of the moving
crack, thus lowering the velocity of the tip. After some time, the daughter crack
stops growing, and thus the crack tip velocity increases, until a new daughter frac-
ture starts to grow, and so on. They also observed that the branching angle for a
longer daughter fracture was smaller than that of the shorter daughter fractures.
These features are all produced by the dynamic lattice models of Heino and Kaski
(1996,1997) described above. The computations of Marder and Liu (1993) for a
perfect crystal, that were described and discussed above, also agree with these data.
As already mentioned, oscillatory fracture patterns were also observed in the
experiments of Yuse and Sano (1993). They imposed a temperature gradient along
a thin glass plate, from a hot region to a cold one. A microcrack was introduced in
the glass, and the glass was pushed. As the plate started to move the crack jumped
ahead of the thermal gradient and stayed there. It was observed that if the plate
moves slowly, the growing crack remains straight and stable. However, increasing
the velocity to a critical value vc gives rise to a transition whereby the fracture
path begins to oscillate and an instability appears. At still higher velocities crack
branching appears; see Figure 6.6. Ronsin et al. (1995) also provided experimental
data for brittle fracture propagation in thin glass strips, using a thermally induced
stress field. In their experiments the temperature field was controlled by the width
w of the plate, and induced thermal expansion in the sample. It was observed that
for widths below a critical value wc no fracture was formed. For wc < w < wo ,
where wo is a second critical width for the onset of oscillatory cracks, straight frac-
tures were formed and propagated with a constant speed. For w > wo oscillatory
fractures were generated that became more irregular as w was increased beyond wo .
Figure 8.27. Schematics of fracture pattern on impact on a thin glass plate. Hertzian fracture
is at the center (after Åström and Timonen, 1997b).
phenomenon very nicely. In his experiments thin square-shaped glass plates were
supported at the edges, and a small and heavy ball was dropped on them from a
point above the plates’ centers. The resulting fracture patterns consisted of three
types of cracks. If the impact velocity was high, only a circular hole at the point of
impact was formed, which is usually referred to as a it Hertzian fracture. However,
at lower impact velocities, radial and tangential cracks also appeared. An example
is shown in Figure 8.27. The radial cracks were fairly straight and directed outwards
from the point of impact, while the tangential ones formed a more or less circularly-
symmetric crack. At still lower impact velocities, only radial cracks were formed,
while if the impact velocity was too low, no crack was formed at all.
A simple and interesting model of dynamic fracture, in which the material cracks
as a result of an impact, was proposed by Åström and Timonen (1997b). They used
a triangular lattice of beams. The dynamics of the system were calculated using a
discrete form of Newton’s equations of motion given by
$ % $ % $ %
1 1 2 1 1
M + D u(t + t) = M − C u(t) − M − D u(t − t),
t 2 2t t 2 t 2 2t
(172)
where M is a diagonal mass matrix, C is the stiffness matrix, D is a diagonal
damping matrix, u is the vector that contains the displacements of the sites from
their equilibrium positions, and t is the length of the discrete time step. The
stiffness matrix that was used was that of slender beams (that is, a beam in which
bending is much larger than shear deformation). The boundary conditions imposed
on the lattice in the xy plane were such that the sites at the boundaries lattice are
constrained to remain at their original positions, while the sites in a circular area in
the middle of the lattice were forced to move a distance −vt in the z direction. For
the lattice in the xz plane only the sites at its left and right edges were constrained to
8.4. Fracture of a Brittle Material by an Impact 443
Figure 8.28. Radial fractures in a lattice with an externally applied in-plane strain. The
results, from top left corner and clockwise, are for times t = 200, 350, 700 and 500 (after
Åström and Timonen, 1997b).
remain at their original positions, while a number of sites in the middle of the upper
boundary moved in the negative z-direction. A beam was considered as broken if
its elongation exceeded its pre-assigned threshold. However, beam fracture was
not instantaneous. It was assumed that, at the threshold, the Young’s modulus of a
beam begins to decrease linearly in time until it reaches zero. The rate at which the
modulus decreased was expressed by a parameter r = c/t, where c is a constant.
Figure 8.28 presents the crack patterns in a lattice with an externally applied
in-plane strain: r = 0.1/t, v = 1/600t, and the in-plane strain, xy = 0.167.
The numbers refer to the time steps at which the snapshots were taken. All the
qualitative features of the experimental crack patterns obtained by Shinkai (1994)
were reproduced by this rather simple model, and therefore it may be used for
studying other aspects of fracture of brittle materials by an impact.
444 8. Brittle Fracture: The Discrete Approach
Figure 8.29. Fracture pattern in a viscoelastic material (after Van Damme et al., 1987a).
An initial microfracture is inserted into the network at its center. Three different
ways of breaking the bonds were considered. In model I only those bonds that
join pairs of sites that are both on the fracture perimeter were broken. In model II
any bond associated with a damaged node—one that had five or fewer unbroken
bonds—was broken, while in model III any of the bonds associated with any of the
sites at the fracture perimeter was broken. The probability of breaking any bond i
is given by
(i − 0 )η
p b = N , (176)
j =1 (j − 0 )
η
where i is the length of bond i, 0 is its equilibrium length, and N is the number
of the unbroken bonds on the perimeter of the fracture. The network was initially
diluted by a small amount to prevent unwanted nonlinearities, and a constant force
was applied on the boundaries of the network. The model produced fractal fracture
patterns with a fractal dimension Df that depended on the parameter η and the
boundary conditions at the perimeter of the growing fracture. The fracture patterns
were quite similar to diffusion-limited aggregates (see Chapters 1 and 5). Other
types of boundary conditions were also used in this model. For example, shear
strain and uniaxial tension were both used (Hinrichsen et al., 1989) in which case
the fracture pattern had an X-like shape. If only bonds in tension were allowed to
break, then only one arm of the X-shape grew.
where a is the activation volume (in 3D) or surface (in 2D) of the system. In a
network model a σ is replaced with La F , where La is the activation length of the
448 8. Brittle Fracture: The Discrete Approach
bonds. If all the bonds are equivalent, then in the limit T → 0 the bond with the
largest strain will always break first, and thus in this limit the probabilistic models
reduce to the deterministic ones described earlier in this chapter. In the opposite
limit, T → ∞, the bond breaking process becomes completely random and thus
represents a percolation process. The process time is incorporated into this model
by the algorithm described above using Eq. (175).
Aside from the early work of Dobrodumov and El’yashevich (1973), the first
model of this type was developed by Termonia and Meakin (1986). In their 2D
model the probability pi that a bond i breaks is given by
pi ∝ exp[αi ei2 /(2kB T )], (178)
where αi is the elastic constant of bond i, and ei is its elongation. Fractal fracture
patterns were generated by this model with a fractal dimension Df 1.3 in 2D.
Termonia and Smith (1986) and Termonia et al. (1985, 1986) developed prob-
abilistic models of mechanical and fracture properties of polymer fibers, which
possess a very complex morphology. In their models there is a distinction between
the primary bonds—those that are parallel to the fiber axis—and the secondary
bonds—those that are perpendicular to the fiber axis. The primary bonds are strong
covalent bonds, while the secondary bonds are the much weaker van der Waals and
hydrogen bonds. Defects are also included in these models by removing a fraction
of the bonds before deformation of the system is started. The bonds break with a
probability given by Eq. (177), but the activation energies and volumes for the pri-
mary bonds were about 2 orders of magnitude larger than those for the secondary
bonds, while p0 was assumed to be the same for both types of bonds. The sec-
ondary bonds were allowed to reform between adjacent sites if their coordinates in
the direction of the primary bonds became equal, whereas primary bonds were not
allowed to do so. Figure 8.30 shows the fracture patterns generated by this model
(under isothermal condition), which are in agreement with typical experimental
observations.
Termonia and Smith (1987,1988) developed models of polymer deformation
and failure in which the effect of chain slippage and the release of entanglements
Figure 8.30. Fracture pattern in the model of Termonia et al. (1985,1986) in which the
probability of failure for the primary and secondary bonds were not the same. The top
pattern corresponds to 150% strain, while the bottom one is for 300% strain.
8.7. Fracture of Thin Solid Films 449
was taken into account. Such effects play a prominent role in failure of polymers.
In their model the rate of failure of van der Waals bonds and that of chain slippage
were both given by Eq. (177), but with different activation energies and volumes.
For the chain slippage process, σ represents the stress difference between two parts
of a chain separated by an entanglement point. Their model used a 2D diamond-
like lattice with nodes that represented the entanglement points between pairs
of polymers. The lattice was then decorated randomly with polymer molecules
that intersect only at the entanglement points until there is an entanglement point
associated with every node. The stress σ is predicted by the classical theory of
rubber elasticity to be
$ %
−1 R
σ = βkB T L , (179)
nc
when nc was the number of chain segments of length between a pair of entan-
glement points separated by a distance R, L(z) = coth(z) − 1/z is the Langevin
function, and β is given by
√
N c nc
β= , (180)
3
with Nc being the number of chain strands per unit volume. The predicted fracture
patterns were found to be in good agreement with experimental observations. More
details can be found in the review by Meakin (1990).
Figure 8.31. A typical fracture pattern in thin films, for which 5000 bonds have been broken
(after Meakin, 1987).
sheets of glass and allowing the water to evaporate slowly along one edge of the gap
between the two sheets. The system resembles a thin film. In the model of Skjeltrop
and Meakin (1988) the points of attachment of the network’s nodes to the substrate
were allowed to move. The force that a weak bond that connects a node i of the
triangular network to the substrate exerts on i is given by F = αw (Ri0 − Ri ), where
αw is the elastic constant of the weak bond, Ri is the position of the ith node, and
Ri0 is its initial position. If |Ri0 − Ri | exceeds a threshold, the point of attachment
to the substrate at Ri0 is moved towards the current position Ri of the node until
|Ri0 − Ri | becomes equal to the threshold. The resulting fracture patterns were in
very good agreement with the experimental patterns.
More recent efforts in this area include those of Crosby and Bradley (1997) and
Leung and Néda (2000). The latter group used a model in which the grains in the
thin films were represented by an array of blocks. Each pair of the neighboring
blocks was connected by a bundle of coil springs having a unit spring constant
and an equilibrium length . Initially, the blocks are randomly displaced by r
about their mean positions on a triangular lattice. For slow cracking on a frictional
substrate, the motion of the grains is over-damped. Thus, the system evolves quasi-
statically with a driving rate which is much slower than the relaxation rate. This
implies that one does not have to solve the equation of motion, but rather update
the configuration of the system according to certain criteria.
In a typical cracking of thin films, the film often hardens and/or weakens in
time, hence tending to contract which is, however, resisted by friction from the
substrate. As a result, as mentioned above, stress is built up and relaxed slowly.
To incorporate such effects in any reasonable model, one can pre-strain the array
of the blocks. Then, two force thresholds, Fs for slipping and Fc for cracking,
are decreased systematically to induce these two modes of motion. For example,
8.7. Fracture of Thin Solid Films 451
Summary
Lattice models offer a detailed account of most aspects of fracture, including
those that had been considered too complex to quantitatively describe, or those
that had simply been ignored. At the same time, there is no conflict between the
lattice models and conventional fracture mechanics. In addition, lattice models
of crystalline materials make the additional prediction that a forbidden band of
velocities exists for cracks. This means that a fracture can only propagate stably
above a finite minimum velocity. Molecular dynamics simulations (see Chapter 9)
of crystalline materials indicate that this forbidden band of states may disappear
at room temperature, but should be observable in low-temperature experiments.
Through precise experiments (most of which were described in Chapters 6 and
7), large-scale simulations of lattice models of quasi-static and dynamic fracture
described in this chapter, and molecular dynamics simulations to be described
in Chapter 9, we now have a deeper understanding of fracture propagation, and
in particular the structure and dynamics of mechanisms of dissipation within the
near vicinity of the tip of a propagating fracture in a brittle material. We now know
that fracture in brittle materials is governed by a dynamic instability that leads
to repeated attempts for fracture branching. Although the instability also appears
in dynamic finite-element simulations of fracture (see Chapter 7), it appears to
have no analytical explanation in a continuum framework. In fact, many classical
models of the cohesive zone have been shown to be ill-posed in that, they admit a
set of possible states under identical conditions. Theories formulated on a lattice,
on the other hand, do not exhibit such difficulties. It remains to be seen whether
a simple continuum limit exists, or whether a crucial ingredient in understanding
fracture is the discreteness of the underlying atoms.
However, despite considerable progress, our understanding of fracture propaga-
tion phenomena is still incomplete. In addition to not having a simple continuum
limit of the discrete models, we must keep in mind that the most detailed and precise
experiments on brittle fracture have so far been carried out in amorphous materials
at room temperature, whereas the most detailed theories currently available apply
mostly to very low (or zero) temperatures. A general theoretical framework for
analyzing fracture propagation over a wide range of brittle materials has not yet
been developed.
Part III
9.0 Introduction
In all the chapters of Volume I, as well as in the present Volume so far, we have uti-
lized continuum mechanics and lattice models to describe modeling and simulation
of morphology of heterogeneous materials and estimating their effective proper-
ties. These models are appropriate for microscopic as well as macroscopic length
scales, but cannot provide any insight into materials’ properties at the smallest
length scales, namely, the molecular scale. In this chapter we describe and discuss
modeling and simulation of materials and their properties at the molecular scale.
To achieve our goal we describe three important theoretical and computational
tools that have been developed over the past three decades, namely, the density
functional theory (DFT) and its variants, classical molecular dynamics (MD) simu-
lation, and quantum MD (QMD) technique. The advent of very fast computers and
development of massively-parallel computational strategies have made it feasible
to carry out large scale calculations at the molecular level, and in this endeavor
these three methods have become indispensable tools for predicting the properties
of materials at such length scales.
Prediction of electronic properties and morphology of a material requires, in
principle, quantum-mechanical computation of the total energy of the system and
minimization of this energy with respect to the electronic and nuclear coordinates.
To carry out such computations, one must start with the Hamiltonian of the system
which, for a system of N electrons and N nuclei with charges Zn , is given by
N
pi2 Pn2
N
1 1
N
N
e
H= + +
2m 2Mn 2 4π ε0 |ri − rj |
i=1 n=1 i=1,i=j j =1
N
N
N
N
1 Zn e 2 1 1 Zn Zn e2
− + , (1)
4π ε0 |ri − Rn | 2 4π ε0
|Rn − Rn |
n=1 i=1 n=1,n=n n
where pi and Pn are the momenta of the ith electron and the nth nucleus, re-
spectively (subscript i refers to the electrons and n to the nuclei), ri and Rn are
their position vectors, m is the electron mass, Mn is the mass of the nth nucleus,
e is the electron charge, and ε0 is the permittivity. The first two terms of Eq. (1)
represent the kinetic energy of the electrons and nuclei, respectively, while the
third and fourth terms are, respectively, the result of Coulomb repulsion between
456 9. Atomistic Modeling of Materials
the electrons and Coulomb attraction between electrons and nuclei. Equation (1)
is too complex for use in exact computations, especially when one must deal with
a large system, and therefore reasonable approximations must be made in order to
make the computations feasible. One obvious simplification can be made by taking
advantage of the fact that there is a large difference in mass between the electrons
and nuclei, while the forces on the particles are the same. Therefore, the elec-
trons respond essentially instantaneously to the motion of the nuclei. As a result,
electronic and nuclear coordinates in the many-body wave function can be sepa-
rated, and the nuclei can be treated adiabatically. This separation is the well-known
Born–Oppenheimer approximation which reduces the solution of the many-body
problem to that of the dynamics of the electrons in some frozen-in configura-
tions of the nuclei. Thus, the Hamiltonian of the system in the Born–Oppenheimer
approximation is given by
N
pi2 1 1
N
N
e2 1 Zn e 2
N N
H= + − . (2)
2m 2 4π ε0 |ri − rj | 4π ε0 |ri − Rn |
i=1 i=1,i=j j =1 n=1 i=1
where ∗ denotes
4 a complex
4 conjugate, and 4 the following notation convention has
been used: dx = s d 3 r ; that is, dx denotes a sum over the spin s and
an integral over the spatial coordinate r . We have used the standard atomic units
(using the Bohr radius, and the electron mass and charge as the basic units) so
that, for example, instead of having the usual factor h̄2 /(2m)∇ 2 we have 12 ∇ 2 . We
use this convention throughout this chapter. Equations (3) and (4) constitute the
well-known Hartree–Fock theory, also known as the self-consistent-field theory.
Note that the fourth term of Eq. (4) is a non-local term, since although the operator
acts on k , its value at r is determined by the value assumed by k at all possible
positions r . Note also that the ground state electron density ρ(r) is given by
ρ(r) = |i (r)|2 . (5)
i
9.0. Introduction 457
scales, fracture itself is the severing of the bonds at atomic scale. This im-
plies that MD simulation of dynamic fracture in materials that account for
the phenomena at molecular scales does not have to be very large, although
large-scale MD simulations of dynamic fracture with up to about 108 atoms
have been carried out (Abraham et al., 1997a,b; see below).
Our discussions in this chapter are also a prelude to Chapter 10 where we
describe how atomistic and molecular modeling of materials is integrated with the
microscopic and macroscopic methods of the previous chapters in order to develop
a multiscale approach for investigating materials’ properties over several widely
disparate length scales.
TDDFT. However, even within this extension, certain problems in solids persisted
for some time, such as the wrong Kohn-Sham band gap, regardless of the type of
approximation used for EXc . Tokatly and Pankratov (2001) showed that, at excita-
tion frequencies, EXc (r, r ) exhibits a highly non-local behavior with a range that
is as large as the system itself, and hence it diverges as the system’s size becomes
very large. They developed a perturbation technique which maintains a correct
electron density in every order of the perturbation theory, and therefore removed
this unphysical feature of the TDDFT.
However, in general, the DFT has been designed for predicting the ground-state
properties, and the Kohn–Sham eigenvalues are actually the derivatives of the
total energy with respect to the occupation numbers of these states (Janak, 1978).
Therefore, it may be appropriate to interpret k and k as only auxiliary variables
that are used for constructing the ground-state energy and density, because there
are many materials for which interpretation of k as the excitation energy is wrong
and leads to erroneous results. Examples include band gaps in semiconductors and
insulators. This should not be surprising because the DFT scheme has really been
designed for computing the ground states only.
Alternatively, one can use an indirect method for carrying out the minimization
of the total energy functional. Among the most successful such indirect methods
is the QMD method of Car and Parrinello (1985). We will describe this method in
Section 9.3 after describing the classical MD technique.
are rather badly behaved, and therefore one must be careful in using a gradient
expansion, an issue that will be discussed in the next section. Such difficulties
provided the prime motivation for developing the local-density approximation
which ignores all corrections to the exchange-correlation energy at any point r
due to heterogeneities in its vicinity. Instead, one makes the ansatz that
EXc [ρ(r)] = Xc [ρ(r)]ρ(r) d 3 r, (18)
where rs is called the Seitz radius (such that ρ = 3/4π rs3 ), ζ = (ρ↑ − ρ↓ )/ρ,
and t = |∇ρ|/(2ks φρ) is a dimensionless density gradient. Here, φ(ζ ) = 12 [(1 +
ζ )2/3 + (1 − ζ )2/3 ], and ks is the Thomas-Fermi screening wave number.
Moreover, c (rs , ζ ) = (e2 /a0 )φ 3 [γ ln(rs /a0 ) − ω], with γ = (1 − ln 2)/π 2
0.031091 and ω 0.046644.
The function H is selected in such a way that EXc satisfies several rigorous
constraints. The proposed form for H is given by
$ 2% $ %
e β 1 + At 2
H = γ φ3 ln 1 + t 2 , (23)
a0 γ 1 + At 2 + A2 t 4
with
$ % −1
β a0 c
A= exp − 3 2 − 1 . (24)
γ γφ e
In these equations, e is the electron charge, a0 = h̄2 /me2 , and β 0.066725.
The EX portion of EXc is written as
EX = ρ(r)x (ρ)FX (s)d 3 r, (25)
where X (ρ) = −3e2 (3π 2 ρ)1/3 /(4π ), and s = (rs /a0 )1/2 φt/c is, similar to t, a
dimensionless density with c 1.2277. The function FX (s) is given by
κ
FX (s) = 1 + κ − , (26)
1 + µs 2 /κ
with κ = 0.804 and µ = βπ 2 /3 0.21951. Let us emphasize that, although we
only provided here the numerical values of the several constants that appear in
these equations, they are in fact related to fundamental physical constants, and
do not represent numerical fits to some experimental data. These equations have
been shown to provide very accurate predictions for atomization energies of many
molecules, and therefore are widely used in the DFT computations.
9.1. Density-Functional Theory 465
where the reciprocal lattice vectors l are defined by, G · l = 2π m for all l, with l
being a lattice vector of the crystal and m an integer. Therefore,
i (r) = ci,k+G exp[i(k + G) · r]. (29)
G
In practice, this series is truncated to include only plane waves that have kinetic
energies less than a cutoff energy. This introduces some error into the computations,
but the error decreases with increasing energy cutoff. A suitable value of the energy
cutoff can be selected by an optimization technique (Rappe and Joannopoulos,
1991), since the value of the cutoff is not included in the theory itself. Substitution
of Eq. (29) in (16) converts the Kohn–Sham equations into a relatively simple
set of equations in terms of the coefficients ci,k+G and the eigenvalues k in
which the kinetic energy is diagonal. Even for systems which contain aperiodicity,
such as those with defects, use of a supercell makes the system amenable to this
type of analysis. In any event, the solution of the Kohn–Sham equations, when
written in terms of the eigenvalues of the coefficients ci,k+G , can be obtained by
diagonalization of the associated matrix. The size of the matrix is dictated by the
choice of the cutoff energy, and can be prohibitively large if the system contains
both valence and core electrons. The problem can be overcome by use of the
pseudo-potential approximation which we now describe.
9.1. Density-Functional Theory 467
It is well known that many properties of solid materials depend on the valence
electrons much more strongly than on the core electrons. This fact is exploited
in the pseudo-potential approximation by which one removes the core electrons
and replaces them and the strong ionic potential by a weaker pseudo-potential
that acts on a set of pseudo-wave functions, rather than the true valence wave
functions (Vanderbilt, 1990). The valence wave functions oscillate strongly in
the region occupied by the core electrons, due to the strong ionic potential in this
region. Ideally, the pseudo-potential must be constructed in a way that its scattering
properties for the pseudo-wave functions are identical to those of the ion and the
core electrons for the valence wave functions. But, this must be done in a way that
the pseudo-wave functions have no radial nodes in the core region, where the total
phase shift produced by the ion and the core electrons will be greater by π , for each
node that the valence functions had in the core region, than the phase shift produced
by the ion and the valence electrons. The phase shift generated by the ion core is
different for each angular momentum component of the valence wave function,
and therefore the scattering from the pseudo-potential must depend on angular
momentum. Outside the core region, the two potentials and their scatterings are
identical. The most general form for a pseudo-potential is then given by
Ups = |lmul lm|, (30)
lm
where |lm are the spherical harmonics, and ul is the pseudo-potential for angular
momentum l.Acting on the electronic wave function with this operator decomposes
the wave function into spherical harmonics, each of which is then multiplied by
the relevant pseudo-potential ul .
A pseudo-potential that does not depend on the angular momentum components
of the wave function is called a local pseudo-potential, which is a function only
of the distance from the nucleus. While it is possible to generate arbitrary, pre-
determined phase shifts for each angular momentum state with a local potential, in
practice there are limits to the amount that the phase shifts can be adjusted for the
different angular momentum states, since one must preserve the smoothness and
weakness of the pseudo-potential, without which it becomes difficult to expand
the wave functions using a reasonable number of plane-wave basis states.
In order for the exchange-correlation energy to be represented accurately, the
pseudo and the true wave functions must be identical outside the core region. If
one adjusts the pseudo-potential to ensure that the integrals of the squared ampli-
tudes of the true and pseudo-wave functions inside the core region are identical,
the equality of the pseudo and true wave functions outside the core region is
guaranteed. Pseudo-potentials that possess this property were first constructed by
Starkloff and Joannopoulos (1977), and have been shown to be highly accurate
for heavy atoms. Moreover, Hamann et al. (1979) pointed out that, a match of
the pesudo and real wave functions outside the core region also ensures that the
first-order energy dependence of the scattering from the ion core is correct, so that
the scattering is accurately described over a wide range of energy. A technique for
468 9. Atomistic Modeling of Materials
where rN = {r1 , r2 · · · , rN } is the position coordinates of all the particles, and r̂ij
is a unit vector along ri − rj , pointing from particle i to j . Then, the equation of
motion for particle i is given by
d 2 ri (t)
mi = Fi (rN ) + Fe , (32)
dt 2
where mi is the atomic mass of particle i, and Fe represents all the external forces
that are imposed on the system, either by nature (for example, the gravitational
force) or by the experimentalist (for example, an external pressure gradient applied
to the system). Molecular dynamics consists of writing Eq. (32) for all the N
particles of the system and integrating them numerically and simultaneously (since
the equations are coupled through the force Fi ). The solution of this set of equations
describes the time evolution of the system. If the forces between the particles
depend only on their relative positions, then the system’s energy and momentum
are automatically conserved.
9.2. Classical Molecular Dynamics Simulation 471
To check whether the system has reached a steady state, if one exists, quantities
such as the kinetic energy are computed and their variations with the time are
monitored. When these quantities no longer vary with the time appreciably, then
attainment of the steady state has been confirmed. The time to reach a steady state
depends on the initial state of the system, and how far it is from its steady state.
Although the MD approach is, in principle, a rigorous method, in practice, and
similar to almost all computational strategies for studying a phenomenon, it is only
an approximate technique. Thus, it should be used with considerable care. Some
of the problems that one must pay particular attention to are as follows.
(1) The interaction potentials between the particles are not, in almost all cases,
known, and therefore one must use approximate expressions for describing
these potentials. In principle, quantum-mechanical calculations can be used
for determining these forces, but such computations can be subject to errors.
Typically, the interaction potentials or forces are written in terms of several
parameters which are determined either by ab initio computations or by fitting
the results to experimental data (see Section 9.7).
At atomic scale, the interactions are either of intra-molecular or inter-
molecular type. We will discuss the intra-molecular interactions separately,
and for now briefly describe the inter-molecular interactions. In many MD
simulations the interaction potential between a pair of particles, the centers
of which are a distance r apart, is represented by the classical Lennard–Jones
(LJ) potential. It was thought for a long time that this potential is too simple to
mimic the behavior of real materials, particularly brittle ones. However, using
what the physics Nobel Laureate R. P. Feynman emphasized nearly 40 years
ago as a guide, the value of the LJ potential is in its universal nature, since
according to Feynman the single most important statement describing our real
world is that, all things are made of atoms, little particles that move around in
perpetual motion, attracting each other when they are a little distance apart,
but repelling upon being squeezed into one another, which is precisely how the
LJ and similar potentials have been constructed for representing real atoms,
fluids, and materials. For use in MD simulation and modeling of many proper-
ties of materials, including their dynamic fracture which is of interest to us in
this chapter (see Section 9.8), a relatively general potential can be constructed,
part of which is based on the classical LJ potential. This potential is given by
(Holian et al., 1991)
⎧
⎪
⎪ σ 12 σ 6
⎪
⎪ 4 − , r < ri
⎨ r r
U (r) = (33)
⎪
⎪ −a1 (rc2 − r 2 )2 + a2 (rc2 − r 2 )3 , ri ≤ r ≤ rc
⎪
⎪
⎩
0, r > rc
also called the collision diameter. Note that σ is not the same as molecular
diameter of the molecules, although the two quantities are usually close to
each other. The 12-6 part of U (r) is the classical LJ potential in which the
r −12 term represents a hard-core or repulsive potential, while the r −6 term is
its attractive part. The second equation (for ri ≤ r ≤ rc ) has been added for
accommodating the difference between brittle and ductile materials. Here ri
is the inflection point in the potential, and
$ %1/2
1 24U LJ (r i )
rc2 = ri2 5 − 5 1 − 9− (r ) , (34)
25 ri ULJ i
5ri2 − rc2
a1 = ULJ (ri ), (35)
8ri3 (rc2 − ri2 )
3ri2 − rc2
a2 = ULJ (ri ). (36)
12ri3 (rc2 − ri2 )2
If the intention is to study the properties of materials via MD simulations
(without studying their fracture), then typically only the 12 − 6 part of the
potential is used. The force F(r) between the particles is then given by, F(r) =
−∇U (r) for r ≤ rc ; of course, F=0 for r > rc . If the system contains solid
walls, then the interactions between the materials’ atoms and those of the walls
must also be taken into account. A well-known potential due to Steele (1973)
has been used in many simulations (although Steel’s potential represents some
sort of a mean-field approximation, as it assumes that the wall is smooth and
structureless):
$ % $ %4
2 σw 10 σw σw4
Uw = 2πρw w σw 2
− − , (37)
5 z z 3(z + 0.61)3
where w and σw are the energy and size parameters that characterize the
interactions between the atoms in the system and those of the walls, z is the
vertical distance from the wall, ρw is the density of the wall’s atoms, and is
the distance between the atomic layers within the wall.
The accuracy of the MD method depends to a large extent on the accuracy of
the interaction potentials used in the simulation. Lennard–Jones type potentials
are too simple to represent complex atoms and molecules. To remedy the
situation, one can fit the size and energy parameters of the LJ potential, so that
certain predictions of the MD simulations fit, in some sense, the experimental
data. For example, these parameters can be estimated from the properties
of fluids at the critical point, liquids at the normal boiling point, or solids
at the melting point. This method has been reasonably accurate for relatively
simple molecules and might, in some cases, provide qualitative insight into the
behavior of materials. However, many materials of technological interest have
strong and specific chemical interactions that cannot be described by simple,
pairwise additive potentials, such as the LJ potential. Thus, more sophisticated
9.2. Classical Molecular Dynamics Simulation 473
This algorithm was first suggested by Stoddard and Ford (1973). The actual number
of interacting particles (i.e., those that are within a sphere of radius rc , centered at
the center of a given particle) is a function of the molecular density. For example, in
the simulation of a typical liquid state using the LJ potential, the computation of a
single pair-interaction requires about 30 − 40 floating-point operations. Therefore,
a complete force calculation requires of the order of 2,000 floating-point operations
per particle, still computationally intensive, but much more efficient than a full N -
body calculation. Let us point out that, the cut-and-shift procedure cannot be used
if electric and gravitational forces are operative in the system, since they decay
only as 1/r. Such cases must be treated separately; see Section 9.2.8.
algorithm, but is far less susceptible to numerical errors, is the leap-frog algorithm,
according to which one computes the velocity of a particle at midpoint between t
and t + t,
$ % $ %
1 1 t
v t + t = v t − t + F[r(t)], (45)
2 2 m
from which the position of the particle is calculated,
$ %
1
r(t + t) = r(t) + t v t + t . (46)
2
In another modification of the Verlet algorithm, the so-called velocity-Verlet
algorithm, one calculates the position and velocity of a particle from
t 2
r(t + t) = r(t) + tv(t) + F[r(t)], (47)
2m
t
v(t + t) = v(t) + {F[r(t + t)] + F[r(t)]}. (48)
2m
This algorithm is most stable with respect to the finite precision arithmetic, and
requires no additional computations in order to calculate the velocities.
An important property of the Verlet algorithm and its leap-frog modification is
that, the energy that one calculates when using these integration methods does not
exhibit any drift in the total energy. This important and desirable stability is due to
the fact that the Verlet algorithm is time-reversible, and therefore does not permit
steady increase or decrease of the energy for periodic systems. An important time
scale in the simulations is the so-called Poincaré time scale, which is the time
after which a system that starts out with a random configuration returns to its
initial configuration. The total time that one can integrate in any MD simulation is,
however, much smaller than the Poincaré time, and therefore there is the possibility
of having an increasing error in the calculated energy as the equations of motion
are integrated for larger times. However, the Verlet algorithm has an additional
property called symplecticity. This property gives rise to conserved quantities,
and in particular Sanz-Serna (1992) showed that, with this property present in
the integration procedure, the discrete analogue of the total energy (in numerical
integrations one can compute only discrete analogues of the properties of interest)
is rigorously conserved, and that the discrete analogue of the total energy deviates
from its continuum (that is, actual) counterpart by an amount which is of the order
of O(t k ), where k is some relatively large integer. Therefore, the Verlet algorithm
and its leap-frog version hold the deviations of the energy bounded. Such desirable
properties are the main reason for the popularity of the Verlet algorithm and its
modifications. In contrast with the Verlet algorithm, the Gear method (Gear, 1971),
which is a predictor-corrector technique, has lost its popularity, despite the fact
that it requires only one force evaluation per time step, because it is not symplectic
and thus can create major problems for those properties of the system that are
supposed to be conserved and bounded.
9.2. Classical Molecular Dynamics Simulation 477
where mi and vi are the mass and magnitude of velocity of particle i, respectively.
The velocity vi of each particle i is then rescaled by α, i.e., vi (t) → αvi (t). The
rescaling of the velocity also necessitates rescaling of the time t, which will be
478 9. Atomistic Modeling of Materials
discussed shortly. Each time this rescaling is done, the actual temperature of the
system changes and gradually approaches the given fixed temperature Tf . This
method can actually be derived by imposing a constant kinetic energy through a
Lagrange multiplier term added to the Lagrangian of the isolated system (Haile
and Gupta, 1983). However, although the rescaling procedure can be derived, due
to certain assumptions used in the derivation, it still represents some sort of an ad
hoc method. This has motivated the development of more systematic methods for
achieving a given temperature and thus being able to perform MD simulations at
constant temperature. One particularly effective method is based on introducing
an extra force acting on the particles. The force is frictional in nature, is assumed
to be proportional to the velocity of the particles, and therefore affects the kinetic
energy of the system (and hence the system’s temperature) in a direct way. Thus,
the equation of motion for the ith particle is written as
d 2 ri dri
m 2
= Fi (R) − Cf (R, dR/dt) , (51)
dt dt
where the friction parameter Cf (R, dR/dt) is assumed to be the same for all the
particles. The sign of Cf depends on whether heat is added to or extracted from
the system; in the former case Cf < 0 while in the latter case Cf > 0. Various
equations have been suggested for Cf . The best-known equation was proposed by
Nosé (1984) and simplified by Hoover (1985), according to which
dCf i mi vi2 − 3N kB Tf
= , (52)
dt f
3
3
2 3
U (rij ) = ULJ (rim,j n ) + UC (rim,j n ) , (57)
m=1 n=1
where rim,j n is the distance between the interacting pairs (site m in molecule i and
site n in molecule j ), ULJ (r) is the-cut-and-shifted LJ potential described above,
and UC (r) is the Coulomb potential given by
qim qj n
UC (r) = . (58)
r
During the MD simulations one must keep track of the coordinates of a CO2
molecules. They can be represented by the vector Vj , with j = 1, 2, · · · , NCO2 ,
where NCO2 is the total number of CO2 molecules. The vector Vj contains three
cartesian coordinates, (rx , ry , rz ), determining the position of the molecule’s cen-
ter, and two coordinates determining its orientation. The orientation of CO2 can be
determined by a unit vector e = (ex , ey , ez ), directed along the axis of the molecule
(where only ex and ey are independent, as ez = 1 − ex − ey ) and the angle that it
makes with the surface of the system’s walls.
9.2. Classical Molecular Dynamics Simulation 481
where subscript 0 indicates that constraints have not been imposed yet. A constraint
is imposed on the system through a Lagrange multiplier λ(t), which is a function
of time since the constraint should hold for all times. For example, for a rigid
molecule that consists of two atoms, we impose the constraint that the distance
between the two particles is always fixed. When such a constraint, which is often
called the bond constraint, is imposed on the system, then the Lagrangian L of the
new system (with the constraint) is given by
t2 7 8
L = L0 − λ(t) [r1 (t) − r2 (t)]2 − 2 dt. (60)
t1
λ(t) is determined by requiring that the solution must satisfy the constraint. We
discuss this shortly.
We are familiar with the concept of the backbone of a disordered medium. The
same concept can be applied to a material at atomic scale. In this case some of
the material’s atoms form the backbone and are fixed by the bond constraints
discussed above, while the remaining atoms are fixed by linear constraints which
we discuss shortly. A good example is provided by a branched polymer in which
the backbone is made of the multiply connected atoms. To identify the backbone of
any molecular structure, some rules of thumb may be useful. For a planar molecular
structure one can consider three non-colinear atoms as a backbone since they satisfy
the bond constraint, while the rest of the atoms in the structure are constrained
linearly. In a 3D molecular structure, four backbone atoms are subjected to six
bond constraints with the remaining ones to a linear vector constraint each. A good
example is provided by the linear molecule CS2 (Thijssen, 1999), the motion of
which is described by five positional degrees of freedom, two of which define the
orientations of the molecules and three define the position of its center of mass.
Without any constraint, the three atoms have nine degrees of freedom, but three
of them are eliminated by the bond constraints, implying that we still have six
degrees of freedom, instead of the required five. The inclusion of the un-needed
degree of freedom adds to the computations and makes them inefficient. A better
procedure is to fix only the distance between the two sulphur atoms by requiring
that, |rS (1) − rS (2) |2 = 2 , and to fix the position of the carbon atom by a linear
482 9. Atomistic Modeling of Materials
It can be shown that the error in the numerical values of λ(t) so obtained is of the
order of O(t 4 ).
Our discussion so far has been restricted to totally rigid molecules. We now
consider partially rigid molecules that consist of rigid clusters that can move with
respect to one another. For this purpose we describe the algorithm due to Rykaert
et al. (1977), Ciccotti et al. (1982), and Rykaert (1985). Their algorithm, which
is known as SHAKE (the author does not know why this name was given to this
algorithm), is formulated based on the notion that the forces that particles experi-
ence are the physical and constraint forces. If M is the number of the constraints,
then the constraints are written as, Ck (R) = 0, with k = 1, 2, · · · , M, where Ck
expresses the functional form of the constraint, e.g., restriction that the distance
between two particles is always fixed. The constraint forces are given by,
M
Fc = λk ∇ i Ck , (70)
k=1
where λk is the Lagrange multiplier to be determined, and subscript i signifies
the fact that the gradient of constraint Ck must be taken with respect to i. Since
we use the Verlet algorithm for integrating the equations of motion, we have the
particles’ positions at times t − t and t which satisfy the constraints imposed on
the system. Thus, an intermediate position r̃i is first calculated for particle i,
r̃i (t + t) = 2ri (t) − ri (t − t) + t 2 Fi [ri (t)], (71)
where Fi represents all the physical forces that the particle experiences. The true
position is then computed from
M
ri (t + t) = r̃i (t + t) − λk ∇ i Ck (rN ), (72)
k=1
(j ) Ck (rN )(o)
λk = . (75)
t 2 { i ∇ i Ck (r )
N (o) · ∇ C [rN (t)]}
i k
484 9. Atomistic Modeling of Materials
The iteration continues until the constraints are satisfied numerically to within a
fixed acceptable error. This algorithm has turned out to be highly efficient and
accurate.
The sum over 1/R is over the locations R of the periodic replicas of the system.
To discuss this further, we consider a concrete example—transport of charged
particles in a disordered material, especially one with a quenched distribution of
charge centers. This problem is relevant to many important phenomena, such as
dynamic response of non-metallic materials, e.g., ionic glasses and polymeric and
glassy conductors, highly defected crystals, and porous materials that are used for
catalytic and separation processes. Although this problem had been studied exten-
sively, until recently no consensus regarding the nature of the transport process had
emerged. In particular, if R 2 (t) is the mean square displacement of the mobile
charged particles at time t, one expects to have
R 2 (t) ∼ t α , (77)
where α = 1 for diffusive transport. However, the precise value of α was a contro-
versial subject with some researchers claiming that α > 1, while others believing
that α ≤ 1, in which case the transport process is anomalous (see Chapter 6 of Vol-
ume I for a discussion of anomalous transport). This controversy also prevented
interpretation of experimental data for diffusion of ions in random media with a
distribution of charge centers. For example, during diffusion of ions through zeo-
lites, which are porous catalytic materials with a distribution of charge center (ions
and cations), it has been observed that, upon changing the charge on the diffusing
particles (i.e., making the disorder stronger), the diffusivity decreases by orders of
9.2. Classical Molecular Dynamics Simulation 485
magnitude. Despite its great importance, no efficient and reliable computer simu-
lation of this problem was carried out for many years, because (1) the Coulombic
interactions between the particles are long-ranged, and (2) the charge centers give
rise to deep potential wells that may capture the mobile particles for long periods
of time and slow down their motion.
Let us now describe how MD simulations of this problem can be carried and, in
particular, how the effect of long-range Coulombic interactions can be taken into
account by an efficient and reliable algorithm developed by Mehrabi and Sahimi
(1999). They used both a continuum and a lattice representation of the system. The
continuum representation was used when the fixed charge centers were distributed
randomly in the medium. The lattice, which was simple-cubic, was utilized when a
potential-potential correlation function, defined below, was utilized for generating
the potentials due to the fixed charge centers. At time t = 0 the charged mobile
particles are distributed randomly in the system, but as they move correlations
develop between them. In addition to the Coulombic interaction, a short-range,
LJ-type repulsive interaction (i.e., ∼ r −12 ; see Section 9.2.1) was also used to
prevent capture of a mobile particle by an immobile one with a charge which has
a sign opposite to that of the mobile particle.
The charge centers are either distributed explicitly throughout the medium, or
are represented by their potential distribution, generated by the potential-potential
correlation function. To make the system neutral, equal numbers of the centers
with opposite charges are inserted in the system, and the same is done with the
mobile particles. The Coulomb potential Ui acting on the ith mobile particle is
written as,
(f m) (mm)
Ui = Ui + Ui , (78)
(f m)
where Ui is due to the interaction between the mobile particle and the fixed
(mm)
centers, while Ui is contributed by the interaction between the mobile particles
(f m)
themselves. Ui can be calculated by two different methods (yielding identical
(f m) (mm)
results). In one method, Ui [and also Ui ] is computed by the multipole
expansion method described below. In the second method, one can use the fact
that diffusion of charged particles in disordered media can be viewed as a transport
process in an external potential field generated by the quenched disorder that
represents the fixed charge centers. Thus, instead of directly distributing the charge
(f m)
centers with a given density ρ(r), Ui is formally represented by the solution
of the Poisson’s equation which, for example in 3D, is given by
q f qm ρ(r )
dr
(f m)
Ui (r) = − (79)
4π ε0 |r − r |
where qf and qm are the charges for the fixed and mobile particles, respectively,
and ε0 is the permittivity. The charge density ρ(r) is represented by its correlation
function χρρ (r) which, in the case of Debye-Hückel statistics, is given by
ρ0 κ 2 e−κ|r|
χρρ (r) = ρ0 δ(r) − , (80)
4π|r|
486 9. Atomistic Modeling of Materials
where ρ0 = ρ(r), and κ −1 is the spatial correlation or the screening length. The
depth of the potential wells in which the mobile particles are captured by the
immobile ones, and thus the radius of influence of the traps, is controlled by κ −1 .
The larger κ −1 , the deeper is the potential well, and thus the larger the time spent
in such traps. In the limit κ −1 → ∞, the trapping times become infinitely large,
and therefore the effective diffusivity is zero. The power spectrum χ̂φφ (ω) for the
potential is calculated from that of the charge density χ̂ρρ (ω), since Eq. (79) is a
convolution integral of the charge density and the Green function for the potential
generated by a single charge particle, and therefore in 3D
$ %
qf qm 2 ρ0
χ̂φφ (ω) = . (81)
ε0 ω (ω2 + κ 2 )
2
Note that the 1D version of Eq. (81) is given by, χ̂φφ = ρ0 (qf qm /ε0 )2 /(ω2 + κ 2 ).
Hence, a realization of the potential
√ √ field is generated as follows. Random numbers,
distributed uniformly in [− 3, 3) (this ensures that their power spectrum is 1,
as it should be), are assigned to the sites of!the system. The resulting array is
then Fourier transformed and multiplied by χ̂φφ (ω), and then inverse Fourier
transformed.
(f m) (mm)
Ui and Ui can also be calculated by a multipole expansion method (Ding
et al., 1992; Mehrabi and Sahimi, 1999). In this method particle i interacts with
the nearby particles through the usual Coulomb potential, and with the far away
particles through their pre-calculated
multipole expansions of the potential. The
total potential U (g) (r) = N j U j (r) produced by a group of N charges is
$ % $ % $ %
q 1 1 1 1 .. 1
U (r) = − P · ∇
(g)
+ Q : ∇∇ − O.∇∇∇ + · · · (82)
r r 2 r 6 r
where q, P, Q, and O are, respectively, the monopole, dipole, quadrupole, and
octapole moments of the group of charges around the origin. In practice, we write
⎛ ⎞
q 1 1
U (g) (r) = + 3 Pα rα + 5 ⎝ Qαα rα rα + Qαβ rα rβ − r 2 Q⎠
r r α 2r α α β
⎡ ⎛ ⎞
1 ⎣ ⎝
+ 7 15 Oααα rα rα rα + Oααβ rα rα rβ + Oαβγ rα rβ rγ ⎠
6r α α β α β γ
−9r 2 Oα + ···, (83)
α
with
q= qi , (84)
i
Pα = qi Riα , (85)
i
Qαα = 2
qi Riα , (86)
i
9.2. Classical Molecular Dynamics Simulation 487
Qαβ = qi Riα Riβ , (87)
i
Q= qi Ri2 , (88)
i
Oα = qi Ri2 Riα , (89)
i
Oααα = 3
qi Riα , (90)
i
Oααβ = 2
qi Riα Riβ , (91)
i
Oαβγ = qi Riα Riβ Riγ , (92)
i
where r = |r|, Ri is the position vector of the ith charge, α, β, and γ stand for the
coordinates x, y, and z, and qi is the charge of the ith (fixed or mobile) particle.
A highly efficient simulation technique is fundamental to this study. Hence,
in addition to taking advantage of the multipole expansion, the 3D simulation
cell is divided into 8 smaller equal boxes, called children of the original box
[see, for example, Greengard and Rokhlin (1987) and references therein]. Each
child box is a parent to 8 smaller boxes, with the division continuing up to a
certain level which is called the maximum level (maxlevel) of division. The data
needed for each particle, i.e., its position and type (mobile or fixed), are stored in
a particle object. A cell object contains a list of its current particles. Each particle
is also “connected” to the next and previous particle in the list. After setting up the
entire data structure, the multipoles of each cell around its center at the maxlevel
are calculated using the above expressions. Then, the multipoles of the parent
cells are computed by translating and adding the multipoles of their children by
a displacement vector = (x , y , z ). In terms of the old quantities, the new
translated (primed) quantities are given by,
Pα = Pα − qα , (93)
Each particle’s potential energy is divided into Uinear and Uifar . A particle in
a cell at the maxlevel interacts with all other particles in the same cell and in
the neighboring cells by the usual Coulomb potential, thus yielding Ujnear . It also
interacts with its parent’s neighbors’ children through the corresponding multipole
expansions. Computations continue up to the entire simulation cell, hence yielding
Uifar . In this way, the number of the cells that interact with each particle is drastically
reduced as one moves away from the particle. For example, in 3D with four levels
of division the number of the interacting cells is only 415, rather than the original
4069 cells. This method is highly efficient for taking into account the effect of
Coulomb and other long-range potentials.
However, even with such an efficient algorithm, MD simulation of this problem
requires intensive computations. If the simulations are not carried out for long
enough times, one may not be able to obtain the true asymptotic (long time)
behavior of the system. As an example, consider the problem in a 1D system,
such as a highly anisotropic material so that the motion of the mobile particles is
restricted essentially to one direction. In this case, the mobile particles can only
travel in the space between themselves, since they cannot “jump” over each other.
There are many relatively fast diffusive jumps in the mean square displacements
of the particles after certain periods of time. In between the jumps one has a slow
motion that causes the overall transport to be anomalous [i.e., α < 1 in Eq. (77)],
not only in 1D but also in 2D and 3D. The mobile particles can be trapped in the
potential wells (traps) that the quenched distribution of the charge centers creates.
The traps have a finite sphere of influence, such that for any particle i within the
sphere the potential difference Ui for a displacement that can take i out of the
sphere is very large, and thus the probability of an appreciable jump is small.
However, after some time the particles are close to the boundary of the traps and
escape with a displacement that takes them out of the traps. They then resume their
diffusive motion until they are captured by another trap, and so on.
number of practical problems. One can use the velocity autocorrelation function,
but this quantity can only be used for predicting the tracer or self-diffusivity Ds
of a species (i.e., when the system is very dilute) via the Green-Kubo equation:
, -
1 ∞
N
Ds = [vi (t) · vi (0)] dt, (100)
3N 0 i=1
the fluxes are calculated at the center of the transport region. The permeability ki
of species i is then calculated from
Ji nLJi
ki = = , (104)
Pi /nL Pi
where Pi is the partial pressure for species i along the pore. The transport diffu-
sivity De is then obtained from the Fick’s law, Ji = −De ∂ρi /∂x, where ∂ρi /∂x is
the adsorption density gradient of component i along the x-direction. Another im-
portant property for the problem that we are describing is the dynamic separation
factor S21 defined as
k2
S21 = . (105)
k1
The NEMD method that we described here has proven to be a practical tool for
simulating transport properties of fluid mixtures in not only a carbon nanopore, but
also in nanoporous materials, such as a variety of membranes, with an intercon-
nected network of nanopores (Xu et al., 2000a). Its predictions for some properties
of interest are in good quantitative agreement with experimental data (Xu et al.,
2000b). Other NEMD methods have been discussed by Rapaport (1995).
Here mψ is a fictitious mass associated with the electronic wave functions, giving
rise to the kinetic energy term of the Lagrangian that arises due to the fictitious
dynamics of the electronic degrees of freedom, E is the Kohn–Sham energy func-
tional, described in Section 9.1, which plays the role of the potential energy, RI
is the position of the ion I , n defines the size and shape of the periodic unit cell,
and ˙ i = di /dt. As Eq. (106) suggests, the details of the kinetic energy do not
9.4. Quantum Molecular Dynamics Simulation: The Car–Parrinello Method 493
matter. What is more important is that the mass m should be much smaller than
the nuclear masses which would prevent transfer of energy from the classical to
the quantum degrees of freedom over long periods of the numerical simulations.
The electronic wave functions are subjected to the orthonormality constraints:
i∗ (r)j (r) d 3 r = δij , (107)
i (108)
+ λij i∗ (r)j (r) d r − δij .
3
i j
Mathematically, the Lagrange multipliers λjj ensure that the wave functions re-
main normalized, while λij (with i = j ) guarantee that the wave functions remain
orthogonal. Physically, the Lagrange multipliers can be thought of as providing ad-
ditional forces acting on the wave functions, ensuring that, at any given time t, they
remain orthonormal as they propagate along their MD paths. The Lagrange mul-
tipliers are symmetric, λij = λj i , and represent 12 N (N + 1) independent values
which are determined by the 12 N (N + 1) orthonormality conditions. The iterative
algorithm SHAKE, described in Section 9.2.7, can be used for determining the
Lagrange multipliers, and was in fact utilized by Car and Parrinello.
makes the computations tractable, it also creates a new problem: At the end of each
time step the wave functions will not be exactly orthonormal, so that a separate
orthonormalization must also be carried out at the end of each time step (see
below).
However, since a separate orthonormalization step must be carried out at the end
of each time step, one may remove the orthogonality constraints from the equations
of motion and use partially constrained equations of motion. After these equations
have been integrated, the orthogonality constraints are imposed again, and the
Lagrange multipliers λij for the normalization constraints are approximated by
the expectation values of the energies of the states given by
λi = i |H|i . (111)
With this approximation the equations of motion for the wave functions are given
by
d 2 i
m = −(H − λi )i , (112)
dt 2
which ensures that the acceleration of an electronic state is always orthogonal to
that state, and that the acceleration becomes zero when the wave function is an
exact eigenstate. However, more generally, one can start from the estimate
˙ i |
λij = 2i |H|j − m ˙ j , (113)
and proceed with integration of Eq. (110). The difference between (111) and (113),
aside from the approximate nature of (113), is that the latter depends on m which
itself must be somehow selected carefully.
where the superscripts n and o refer to the new and old estimates of the wave
functions, respectively. If Eq. (117) is applied repeatedly to the old estimates, then
the electronic wave functions can be made orthogonal to any desired accuracy. For
496 9. Atomistic Modeling of Materials
example, if algorithm (117) is applied to two wave functions, they will be exactly
orthogonal after only one application of the algorithm. Moreover, if the estimates
of the wave functions are orthonormal up to order t 4 (the accuracy of the Verlet
algorithm), then over a given time step algorithm (117) changes them to an extent
within the same order. In general, the number of times that algorithm (117) must
be applied depends on the number of the wave functions to be computed and their
initial departure from orthogonality. However, algorithm (117) does not preserve
normalization of the wave functions, and therefore they must be normalized after
each application of the algorithm.
The Kohn–Sham energy functional is minimized by any set of wave func-
tions that are a linear combination of its lowest-energy eigenstates. Under the
MD scheme, the wave functions that are obtained after orthogonalization will be
stationary, implying that in the MD method each wave function will converge to a
linear combination of the lowest-energy Kohn–Sham eigenstates. This can create
severe problems for treating metals, since the ability to converge to Kohn–Sham
eigenstates (and not to their linear combinations) is highly important for metal-
lic materials. Several methods have been proposed for addressing this problem
(see, for example, Pederson and Jackson, 1991, and references therein). The ac-
tual Kohn–Sham eigenvalues can be found by either diagonalization of the matrix
A = (Aij ), the entries of which are given by
Aij = i |H|j , (118)
or by the Gram–Schmidt algorithm (see, for example, Noble and Daniel, 1977)
which is similar to (117) but without the factor 1/2 in front of the sum. Thus, instead
of using algorithm (117) which generates wave functions that are linear combina-
tions of the Kohn–Sham eigenstates, one can use the Gram–Schmidt method which
results in orthogonal wave functions in such a way that all of the higher-energy
wave functions are forced to be orthogonal to the lowest-energy wave functions.
This in turn forces each state to converge to its lowest possible energy while sat-
isfying the constraint that it be orthogonal to all states below it. Therefore, the
set of lowest possible single-particle levels under these constraints comprises the
Kohn–Sham eigenstates.
on the ions is not a physical force, rather it is a force that the ions experience
from a particular electronic configuration. However, it is not difficult to show that,
when the electronic wave functions are the eigenstates of the Hamiltonian, then
the second and third terms of Eq. (123) vanish, and therefore, in this case, ∂E/∂RI
yields the true physical force on the ions. This important result is known as the
Hellmann–Feynman theorem (Hellmann, 1937; Feynman, 1939), and in fact holds
for any derivative of the total energy.
The Hellmann–Feynman theorem greatly simplifies the task of computing the
physical forces acting on the ions and the integrated stresses that are exerted on
the unit cell, since it allows one to compute these quantities only when the wave
functions are very close to exact eigenstates. Once the forces and stresses have been
computed, the positions and the shape and size of the unit cell are also calculated
using Eqs. (120) and (121), which represent their equations of motion. Each time
the positions of the ions or the shape and size of the unit cell change, the electrons
must be close to the ground state of the new ionic configuration in order to compute
forces and stresses for the new configuration.
The simplest way that the Hellmann–Feynman forces are used is for determining
the position of a local energy minimum of the ionic system. In this scheme the ions
are moved along the directions of the Hellmann–Feynman forces until the residual
forces (i.e., the deviations from ∂E/∂RI ) on all the atoms are smaller than a given
value. These forces cause the ions to fluctuate around their equilibrium positions.
The residual forces acting on the ions are never zero. If the system is to approach
the minimum energy state, the magnitudes of the errors in the Hellmann–Feynman
forces must be reduced, implying that the electronic configuration must be relaxed
closer and closer to the instantaneous ground states as the ionic configuration
approaches the local energy minimum.
plane-wave basis states constant results in a change in the cutoff energy for the
basis set. If we increase the number of plane-wave basis states by increasing the
cutoff energy for the basis set, it will eventually result in a reduction in the total
energy of the system. However, if the cutoff energy is large enough to achieve
absolute convergence, the change in the total energy will be zero. In practice, most
pseudo-potential computations are carried out with a cutoff energy at which energy
differences have converged, but at which the total energies have not converged,
in which case the diagonal components of the Pulay stresses acting on the unit
cell will have non-zero values. It can be shown that the change in the total energy
per atom is independent of the details of the ionic configuration, provided that the
cutoff energy for the basis set is large enough for the energy differences to have
converged. This facilitates computation of the Pulay stresses, since they can be
calculated once and for all from the change in the total energy of a small unit cell
as the cutoff energy is varied.
where u0α is the local pseudopotential for the l = 0 angular momentum state
[see Eq. (30)]. Integral (126) is zero outside the core region since, by defini-
tion, the potentials are identical outside this region. This non-Coulomb part of the
pseudo-potential
does contribute to the total energy at G = 0, which is given by
N −1
c α N α u α,core , where N is the total number of electrons in the system, Nα
500 9. Atomistic Modeling of Materials
is the total number of ions of species α, and c is the volume of the unit cell. We
remind the reader that the Coulomb energy of the ionic system can be calculated
by the method described in Section 9.2.8.
(3) The equations of motion for the electronic states are integrated using the Ver-
let algorithm, and the resulting wave functions are orthogonalized, either by
algorithm (117) or by the Gram–Schmidt method; the wave functions are also
normalized.
There are several “tricks” that can speed up the computations. For exam-
ple, if we assume a local pseudo-potential approximation and use Eq. (29) in
Eq. (16), then it is straightforward to show that the governing equations for
the coefficients ci,k+G are given by
d 2 ci,k+G
= −ωi,k+G
2
ci,k+G − Bi,k+G , (127)
dt 2
where ωi,k+G and Bi,k+G are two sets of coefficients that arise when Eq. (29)
is inserted into Eq. (16). However, we recognize Eq. (127) as the oscillator
equation, which means that if the Verlet algorithm is to be used for integrating
this equation, the time step must be such that t ωi,k+G < 1 for all of the
plane-wave basis states. This implies that t is restricted by the plane-wave
basis states that have the largest kinetic energies. However, as we discussed
in Section 9.1.4, the largest kinetic-energy basis states contribute the least to
the solution of i , Eq. (29), and this creates an unsatisfactory situation. As
discussed by Payne et al. (1992), a way around this problem is assuming that
the coefficients Bi,k+G are constant over a time step of duration t, in which
case Eq. (127) can be easily integrated analytically over the time step, with
the solution given by
(6) The Kohn–Sham energy functional is minimized; its value gives the total
energy of the system. If the plane-wave basis set, Eq. (29), has been used (which
is always the case if pseudo-potential approximation is used), convergence
tests must be performed to ensure that the calculated total energy has converged
both as a function of the number of terms included in the set and the value
of the cutoff energy that has been used for truncating Eq. (29) after a finite
number of terms.
(7) Integration of the equations of motion for the ionic positions and the coor-
dinates of the unit cell is also done along the QMD computations for the
electronic configurations. It is sensible in the QMD methods to treat the
electronic and ionic systems independently when relaxing the ions to their
equilibrium positions, and therefore it is also possible to use different time
steps for the two systems. As the integration proceeds, the time step for the
ionic system must be progressively reduced as the ionic configuration ap-
proaches the local energy minimum. Such a procedure allows the electronic
configuration to relax closer to its instantaneous ground-state configuration as
the ions approach their equilibrium positions, hence ensuring that the errors
in the Hellmann–Feynman forces are always less than the actual forces acting
on the ions.
The QMD technique allows one to search large regions of configuration space
and locate the deeper energy minima in a very efficient manner. Since the QMD
combines a MD method with the DFT, it makes it possible to study temperature-
dependent effects by a method that is free of the common assumptions about the
nature of the interatomic forces. Figure 9.5 presents the pair correlation function
C(r) for both amorphous and liquid Si, calculated by Car and Parrinello, and its
comparison with the experimental data. Given that the only piece of information
that was supplied to the simulator was the volume of the unit cell, the agreement
between the predictions and the data is truly remarkable.
Figure 9.6, adapted from Jones and Gunnarsson (1989), shows the evolution
of the structure of Se5 molecule, starting with an almost linear geometry, and
obtained by the Car–Parrinello method. The effect of the core region was repre-
sented by a pseudo-potential. The time between successive structures is almost
500 time steps of 3.4 × 10−16 seconds each. The last structure shown in the figure
is stable and agrees with the data. The important point to remember is that, the-
oretically, there are many possible structures that correspond to the local energy
minima, and the QMD method of Car and Parrinello can find the true structure
very efficiently.
Let us mention three different and successful applications of the QMD and
pseudo-potential methods that we just described. These applications represent only
a sample of a very large number of computations that have been reported so far.
Marks et al. (1996) reported the results of ab initio simulations of tetrahedral
amorphous carbon based on the Car–Parrinello method. The simulated structure
was in good agreement with the experimental data. Pickard et al. (2000) reported
the results of computations for a variety of lanthanide- and actinide-containing
compounds. The simulated structures and the associated structural parameters were
in excellent agreement with the experimental data. More significantly, they showed
that the pseudo-potential formulation allows a steady march through the Periodic
Table, in the sense of calculating reliably the structural parameters of compounds
that contain f electrons.
Yoon et al. (2001) employed ab initio pseudo-potential DFT with a linear com-
bination of atomic orbits and an exchange-correlation functional in the generalized
gradient approximation, described in Section 9.1.2, to study structural deformation
and intertube conductance of crossed carbon nanotube junctions. They reported
good agreement between the results of their simulations and the experimental data.
504 9. Atomistic Modeling of Materials
electronic states may be encountered. Such instabilities will not arise in a direct
method. In this section, we describe two of such methods and their application to
minimization of the Kohn–Sham total-energy functional.
The best choice of b1 and b2 for minimizing f (x) along d1 and d2 is obtained by
differentiating Eq. (131) with respect to b1 and b2 and evaluating the result at x3 .
This procedure yields two equations,
(x1 + b1 d1 + b2 d2 ) · ∇ · d1 = 0, (134)
(x1 + b1 d1 + b2 d2 ) · ∇ · d2 = 0. (135)
However, in order for Eqs. (132) and (133) to be consistent with (134) and (135),
one must have
d1 · ∇ · d2 = d2 · ∇ · d1 = 0, (136)
implying that the directions d1 and d2 must be conjugate to each other, hence the
name conjugate-gradient (CG) method. More generally, the directions di and dj
must be such that
di · ∇ · dj = 0, i = j. (137)
Hence, in the CG method one takes the initial direction to be −∇f (x1 ), and
the subsequent directions are constructed from a linear combination of the new
gradient and the previous direction that minimized f (x). In practice, the new
direction di in the ith iteration is obtained from
di = vi + ωi di−1 , (138)
where
vi · vi
ωi = , (139)
vi−1 · vi−1
and ω1 = 0. Note that vi = −∇f (xi ). It has also been observed that in some cases
a better estimate of ωi is given by
(vi − vi−1 ) · vi
ωi = . (140)
vi−1 · vi−1
In the CG method a function is guaranteed to converge to its true minimum. The
reason is that since minimization along the conjugate directions are independent,
each iteration reduces the dimensionality of the vector space by 1. Thus, one
reaches the point at which the dimensionality of the function space is zero, i.e.,
there are no new directions left along which one can minimize the function, and
therefore the trial vector must be at the position of the true minimum. The number
of the iterations needed to reach the true minimum is therefore at most equal to the
dimensionality of the vector space, although in practice it usually takes far fewer
iterations to converge to the true solution.
between the iterations in order to ensure the conjugacy of the search directions.
The ideal method is one that takes advantage of the efficiency of the CG method
without requiring much more computer memory. This can be achieved by updating
a single band at a time. The steepest-descent direction for a single band is given
by
ζim = −(H − λm m
i )i , (141)
where m denotes the iteration number and i labels the band. Here [see Eq. (111)],
i = i |H|i .
λm m m
(142)
However, since the electronic wave functions must be orthogonal, one must en-
sure that the steepest-descent vector is orthogonal to all the other bands. The
components of the steepest-descent direction vector that ensures this is given by
ζim = ζim − j |ζim j . (143)
j =i
The components of the pre-conditioned conjugate directions dim are now given by
[see Eq. (138)]
dim = ηi m + ωim dim−1 , (147)
with [see Eq. (139)]
ηi m |ζi m
ωim = , (148)
ηi m−1 |ζi m−1
with ωi1 = 0. The resulting conjugate directions will be orthogonal to all the
other bands, except the wave function of the present band. Thus, a further
orthogonalization to the present band is done via
di m = dim − im |dim im , (149)
di m
di m = . (150)
di m |dim 1/2
Having constructed the pre-conditioned conjugate direction, the search for the
minimum energy begins. A vector with the following components
im cos θ + di m sin θ (151)
is a normalized vector orthogonal to all the other bands j (with j = i and θ real),
and thus satisfies the constraints of orthonormality required for the electronic wave
510 9. Atomistic Modeling of Materials
functions. The value of θ that minimizes the Kohn–Sham energy functional is then
required by the CG method. It has been found that the following approximation to
the Kohn–Sham energy,
E(θ ) = E + A cos(2θ ) + B sin(2θ ), (152)
is sufficient for locating its minimum. There are three constants, A, B, and E, and
therefore one needs three data points to evaluate them. One piece of information is
provided by E(θ = 0) which is already known. The second data point is supplied
by the fact that since at θ = 0
∂E
= dim |H|im + im |H|dim = 2Re dim |H|im , (153)
∂θ
and that, in order to determine the components ηim of the steepest-descent vector,
H|im has been determined, computing E (0) = ∂E/∂θ , evaluated at θ = 0 is
cheap. The third data point can be taken to be the value of the Kohn–Sham energy
at a point other than θ = 0. This point should be selected to be far enough from
θ = 0 to avoid round-off errors, yet not so far from θ = 0 that the estimate of
E (0) becomes inaccurate. It has been found that θ = π/300 gives very reliable
results. With these three data points, the three constants are found to be
2E(0) − 2E(π/300) + E (0)
A= , (154)
2[1 − cos(2π/300)]
1
B = E (0), (155)
2
2E(π/300) − E (0) − 2E(0) cos(2π/300)
E = . (156)
2[1 − cos(2π/300)]
We can now determine straightforwardly the value of θmin that minimizes the
Kohn–Sham energy, since the stationary point of Eq. (152),
$ %
1 −1 B
θs = tan , (157)
2 A
that lies in the range (0, π/2) is the required θmin for minimizing the Kohn–Sham
energy. To start the next iteration of the CG method, a new wave function, given
by
im+1 = im cos(θmin ) + dim sin(θmin ), (158)
is utilized. Note that each new wave function generates a new charge density, and
therefore the electronic potentials in the Kohn–Sham Hamiltonian must be updated
before starting the next iteration.
In practice, only a few iterations are performed on any given band before moving
to the next band, since there is little to gain from converging a single band exactly
while there are still errors in the estimates of the remaining bands. Once all the
bands have been updated, the CG iterations are started again on the lowest band.
One can also perform the CG iterations on one band until the total energy changes
9.6. Vectorized and Massively-Parallel Molecular Dynamics Simulation 511
by less than a given fraction of the change of energy in the first CG iteration, and
then start iterating on the next band.
are used in the usual way in order to identify neighbors within a distance rc of each
atom. The combined algorithm is made even more efficient by taking advantage of
Newton’s third law which allows one to compute a force for each pairs of atoms,
instead of once for each atom in the pair. This reduces the necessary searches to
only half the surrounding bins of each atom to form its neighbor list. In this way
an atom j is stored in atom i’s list, but not vice versa, hence halving the number
of force computations that must be done.
Another factor that can increase the efficiency of a vectorized MD algorithm
is careful data and loop structures, without which the computer program cannot
be completely vectorized. For example, Grest et al. (1989) combined neighbor
list/link-cell method to create long lists of pairs of neighboring atoms. At each
time step, they updated the lists to keep only those particle pairs that were within
the cutoff distance rc . They also organized the lists into packets in order to prevent
any atom from appearing twice in the lists. In this way, the force computations for
each packet was completely vectorized, resulting in an algorithm that was about
one order of magnitude faster than an algorithm without such data structures.
Another way of developing highly efficient algorithms for MD simulations is
by taking advantage of the natural parallelism that exists in such simulations when
short-range forces are involved. We now discuss such algorithms in detail.
The reason that MD simulations with short-range forces are amenable to par-
allel computations is that calculations of the forces and updating of the positions
can be done simultaneously for all the atoms. Thus, the main goal is to divide
the force computations evenly among the processors so as to achieve maximum
parallelism. There are at least three ways of achieving this, and what follows is a
brief description of each, which are patterned closely after Plimpton (1995).
with a process or 12 Np positions away, after which each processor has the complete
vector r.
The inverse of the expansion operation is also a useful procedure and is com-
monly called a fold operation. Suppose that each processor has stored new force
values in its copy of f. Therefore, processor Pm needs to know the N/Np values
in fm , where each of the values is summed across all Np processors. This is also
shown in Figure 9.7. Thus, each processor exchanges half the vector with a pro-
cessor that is 12 Np positions away, such that it receives the half that it is a member
of, and sends the half that it is not a member of, and then sums the received values
with its corresponding retained sub-vector. This operation is repeated such that at
each step the length of the exchanged data is halved. Note that the expand and fold
operations are optimal, since each processor carries out log2 Np sends and receives
and exchanges N − N/Np additions in the fold operation. The main disadvantage
of this algorithm is that, it requires O(N ) storage on every processor.
The AD algorithm can be implemented in two different ways. In the first one,
which we refer to as the AD1 algorithm, one assumes that, at the beginning of each
time step, every processor has an updated copy of the vector r, and thus “knows”
the positions of all the N atoms. Then, the step-by-step implementation of the AD1
algorithm is as follows.
(1) This step consists of constructing the neighbor lists for all the pairwise in-
teractions that must be computed in block fm . This is not done at every time
step, but after every few steps. It is more efficient for a processor to inspect
all the N 2 /Np pairs in its fm , if the ratio of the physical size of the system to
rc is small. For large simulations, however, in which four or more bins can be
created in each dimension, it is more efficient for a processor to bin all the N
atoms and then inspect the 27 surrounding bins, the length of each is N/Np .
The overall scaling of this inspection process is about N + N/Np .
(2) The neighbor lists are now utilized for calculating the non-zero matrix entries
in Fm . Since each pairwise interaction force is calculated and the force com-
ponents are summed into fm , Fm is never actually stored as a matrix, hence
saving a lot of computer memory. At the end of this step, each processor knows
the total force fm on each of the N/Np atoms that it owns.
(3) The information in (2) is now used to update the positions and velocities of
the particles (i.e., by using them in the equations of motion and integrating
them over one time step).
9.6. Vectorized and Massively-Parallel Molecular Dynamics Simulation 515
(4) The updated positions of the atoms obtained in the previous step are now shared
among all the Np processors, using the expand operations defined above, in
preparation for the next time step.
This algorithm does not take advantage of Newton’s third law. Algorithm AD2
does use this law and thus reduces the time of the computations by decreasing the
cost of communication between the processors. In order to do this, another matrix
G is used such that Gij = Fij , except that Gij = 0 if i > j and i + j is even,
or when i < j and i + j is odd. Thus, for example, step 1 of AD2 is similar to
that of AD1, except that only half as many neighbor list entries are constructed by
each processor because Gm has only half the non-zero entries of Fm . Similarly,
if the neighbor lists are constructed by binning, then although all the N atoms
must be binned, each processor needs to inspect only half the surrounding bins of
each of its N/Np atoms. In step 2 of the AD2 algorithm, the neighbor lists are
utilized for computing the entries of Gm . For the interaction between atoms i and
j , the forces acting on i and j are summed into both i and j locations of force
vector f, implying that each processor must store a copy of the entire force vector,
as opposed to storing only fm done in the AD1 algorithm. After all the matrix
entries are computed, f is folded across all the Np processors, hence enabling each
processor Pm to have fm , the forces acting on its atoms. Steps 3 and 4 of the AD2
algorithm are the same as that of AD1.
Nβ
!
Nα " # .$ %
& '
! " #
Figure 9.8. The division of the permuted force matrix F among 16 processors√ in the force-
√
decomposition algorithm. Processor 6 is assigned a sub-block F6 with size N/ P × N/ P
√
which, to compute its matrix elements, must know the corresponding N/ P −length pieces
xα and xβ and permuted position vector x (after Plimpton, 1995).
(1) Neighbor lists are constructed. If N is small, construction of the lists is most
efficiently
! done by checking all the N 2 /Np possible pairs in Fm . For large N
the N/ Np atoms in rβ are binned and the 27 surrounding bins of each atom
in rα are checked.
(2) The neighbor lists are utilized for computing the entries of Fm . As they are
computed, the entries are summed into a local copy of fα , and therefore Fm
does not have to be stored in matrix form.
(3) By a fold operation within! each of the rows, processor Pm obtains the total
force fm acting on its N/ Np atoms.
(4) Processor Pm utilizes fm to update the positions of the N/Np atoms in rm .
(5) The processors in row α perform an expand operation on their rm sub-vectors,
so that each of them obtains the entire rα . Similarly, the processors in each
column β perform an expand operation on their rm . It is in this step that using
the permuted force matrix F saves extra communications and hence computer
time.
However, the FD1 algorithm does not take advantage of Newton’s third law, and
therefore calculates each pairwise interaction force twice. In algorithm FD2, this
is avoided by using the modified force matrix G, the construction of which was
discussed for the AD2 algorithm. Then, G is permuted in the same way as F to form
G . Step 1 of FD2 is the same as in FD1 (except that half as many interactions are
stored in the neighbor lists). In step 2, for each ij entry, the computed components
of the force are now summed into two sub-vectors instead of one, such that the
force acting on atom i is summed into fα in the location corresponding to row i,
while the force acting on atom j is summed into fβ in its proper location. Step
!
3 consists of three stages. First, the Np processors in column β fold their local
9.6. Vectorized and Massively-Parallel Molecular Dynamics Simulation 517
copies of fβ , resulting in fm , each element of which is the sum of an entire column
of G . Next, the same type of operations are performed for the row contributions,
resulting in fm , each element of which is the sum of an entire row of G . Finally,
the column and row contributions are subtracted element by element in order to
obtain the total forces fm acting on the atoms that belong to processor Pm , which
can now update the positions and velocities of its atoms. Steps 4 and 5 of FD2 are
identical to those of FD1.
east/west exchanges
1 2 3 up/down exchanges
north/south exchanges
Figure 9.9. The spatial decomposition algo-
rithm. In six data exchanges all atom positions
2
in adjacent boxes in the east/west, north/south,
and up/down directions are communicated
(after Plimpton, 1995).
518 9. Atomistic Modeling of Materials
a distance re of processor 1’s box, which will be all of its atoms if its dimension l
in the east/west direction is less than re . Otherwise, it will be those that are nearest
to processor 1. Then, each processor sends its message in the westward direction
and receives a message from the eastward direction which puts it in its second data
structure. Then, the process is reversed with each processor sending a message
to the east and receiving one from the western processor. If l > re , all needed
atomic positions in the east-west direction are acquired with one such exchange.
If, however, l < re , the east-west steps are repeated, with each processor sending
more needed atomic positions to its adjacent processor. This process is repeated
until each processor knows all atoms’ positions within a distance re of its box
(dashed boxes in Figure 9.9).
The same procedure is then repeated in the north/south direction. The only
difference with the east-west step is that the messages that are now sent to the
adjacent processor contain not only the atoms that belong to the processor in its
first data structure, but also any atom positions in its second data structure that are
needed for the neighboring processor. If, for example, l = re , this has the effect
of sending three boxes worth of atom positions in one message. In the final step,
the process is repeated in the up-down direction in which the atoms’ positions
from an entire plane of boxes are sent in each message. This algorithm has several
advantages, some of which are as follows.
(1) If l ≥ re , all the needed atom positions from the 26 surrounding boxes are
obtained in only 6 data exchanges. Moreover, if the topology of the parallel
machine is that of a simple-cubic lattice, the processors can be mapped onto the
boxes in such a way that all six of these processors will be directly connected
to the center processor, as a result of which the passage of information will be
very efficient.
(2) If l < re , then atom information is needed from more distant boxes, but the
information is obtained by only a few data exchanges, all of which are still
with the 6 immediate neighbor processors.
(3) The amount of data communication is minimized, since every processor
obtains only the atom positions that are within a distance re of its box.
(4) No time is spent for rearranging the data structures. All the newly-arrived
information are placed as contiguous data directly into the processor’s second
data structure. The only time spent is for creating the buffered messages that
must be sent.
(5) Message creation is done quickly. The two data structures are scanned only
once after every few time steps, when the neighbor lists are created, in order
to decide which atom positions to send in each message.
Suppose that box number m is assigned to processor Pm which stores the po-
sitions of its N/Np atoms in rm , the first data structure, and the forces acting on
these atoms in fm . The details of the procedure for implementing the SD algorithm
is as follows.
(1) The positions, velocities, and any other information about atoms that are no
longer inside box m are moved from rm to a message buffer. These atoms are
9.6. Vectorized and Massively-Parallel Molecular Dynamics Simulation 519
then exchanged with the six adjacent processors (see above), during which
processor Pm checks for new atoms that are now inside its box, which are
then added to rm . Then, all atom positions that are within a distance re of
box m are obtained by the communication scheme described above. Since the
different messages are buffered by scanning through the two data structures,
lists of included atoms are constructed, after which the two data structure of
the processor are updated. Then, neighbor lists of the N/Np atoms of the
processor are constructed. If two atoms i and j are in the same box, then the
pair (ij ) is stored once in the neighbor list. If i and j are in different boxes, their
corresponding processors both store them in their respective neighbor lists. If
l < 2re , it is more efficient to find neighbor interactions by checking each atom
inside box m against all the !atoms in both data structures of the processor, an
operation that scales as O( N/Np ). If l > 2re , it is more efficient to construct
the list by binning described above, by mapping all the atoms in both data
structure onto bins of size re , and checking the surrounding bins of each atom
in box m for possible neighbors.
(2) Processor Pm uses the neighbor lists to compute all the forces that act on its
atoms. If both atoms i and j are inside the same box, then the computed force
between them is stored twice in fm , once each for i and j . If the two atoms are
in different boxes, only the force acting on the processor’s own atom is stored.
(3) After computing fm , the atoms’ positions are updated.
(4) The updated positions are communicated to the surrounding processors in
preparation for MD simulation in the next time step.
The cost of these steps scales as the volume of the data exchanged. For example,
for the last step, assuming that we have uniform density in the simulation cell, the
cost is proportional to the physical volume of the shell of thickness re around box
m, which is (l + 2re )3 − l 3 . Thus, three cases must be considered. If l < re , data
from many neighboring boxes must be exchanged and the operation cost scales as
8re3 . If l re , the data in all the 26 surrounding boxes are exchanged and the cost
of operation scales as 27N/Np (note that N/Np is roughly the number of atoms
in volume l 3 ). Finally, if l re , only atoms’ positions near the six faces of box
m will be exchanged, and therefore the cost of the communications scales as the
surface area of the box, namely, 6re (N/Np )2/3 . As before, one can take advantage
of Newton’s third law and make the algorithm still more efficient.
There will be load-balance for the FD algorithms only if Fm and Gm are uni-
formly sparse. Thus, if the atoms are ordered geometrically, these two matrices
will not be uniformly sparse, since geometrical ordering generates a force matrix
with diagonal bands of non-zero entries. The way to achieve uniform sparsity is
to randomly permute the atoms ordering.
Finally, the SD algorithms will be load-balanced only if there is roughly the
same number of atoms both in the boxes and in their surroundings. This will
not be the case if there is non-uniformity in the density of the particles, or if
the physical domain is not a rectangular parallelepiped. Hendrickson and Leland
(1995) developed a method for load-balancing the SD algorithms by partitioning
an irregular domain or a system with non-uniformly-distributed clusters of atoms,
although such algorithms add to the computational cost of the MD simulations.
Table 9.3. CPU seconds/time step for the SD1 algorithm. N is the number of atoms used
in the MD simulations, while Np is the number of the processors of the machine (adapted
from Plimpton, 1995).
Problem Size Cray T3D Intel iPSC/860 Intel Paragon
N Lattice Np = 256 Np = 512 Np = 32 Np = 64 Np = 1024 Np = 1904
5 x 102 5x5x5 0.00432 0.00446 0.0129 0.0106 0.00564 0.00634
5 x 104 20 x 25 x 25 0.0289 0.0167 0.420 0.224 0.0174 0.0125
5 x 106 100 x 100 x 125 1.86 0.994 − − 0.914 0.504
108 250 x 250 x 400 − − − − − 9.11
Paragon with neighbor lists consuming the majority of the space. It is clear that
the algorithm is extremely efficient. However, use of optimized version of the
same algorithm increases its efficiency by a factor 2 − 3. For example, on the Intel
Paragon machine with 1840 processors and an assembler version of the algorithm
(as opposed to, for example, the Fortran version), the CPU seconds/time step was
5.5, a factor of nearly 2 more efficient than what is listed in Table 9.3. Similar
computations were performed for the other algorithms discussed above.
Let us point out that, similar to the classical MD simulations, the electronic
structure calculations and QMD computations can also benefit tremendously from
parallelization. In particular, parallel QMD can be a powerful method for comput-
ing realistic interaction potentials for use in the classical MD simulations (see, for
example, Clarke et al., 1992). This is the subject of the next section.
of states which differ significantly from tetrahedral ground state, or when one must
deal with large deformations. For these reasons, developing accurate representation
of the interaction forces between various atoms has been, for many years, an active
research field. We summarize in this section some of the most significant results
that have emerged over the last two decades.
where ρia is the atomic density of the constituents, then the energy would be a
simple function of the atoms’ positions. Note the difference between ρj and ρja :
Whereas ρja is the contribution to the density from atom j , ρj is the total electron
density at atom j .
As shown by Daw and Baskes (1984), the ground state properties of the solid
can be computed from Eq. (159). For example, consider the case of a perfect,
homonuclear crystal. In this case, Eie = E, U = Uij (rij ), and ρ = ρ a . If ρe is
the equilibrium density, then ρe = m ρ(l m ), where l m are the distances between
neighbors, and the sum is over neighbors. Moreover, one has, for every i, ρi = ρe .
Then, the lattice constant is obtained from the equilibrium condition:
m m m m
1 Um li lj
ρm li lj
+ E (ρ e ) = 0, (161)
2 m lm m
lm
9.7. Interatomic Interaction Potentials 523
where lim is the ith component of the position vector to the mth neighbor, Um =
dU/dr, and ρm = dρ/dr, with the subscript m indicating that the derivatives are
to be evaluated at r = l m .
The elastic constants of the crystal at equilibrium are given by
1
Cij kl = [Bij kl + E (ρe )Wij kl + E (ρe )Aij Akl ], (162)
0
where
1 (Um − Um / l )li lj lk ll
m m m m m
Bij kl = , (163)
2 m (l m )2
/ l m )l m l m l m l l
(ρm − ρm i j k m
Wij kl = , (164)
m
(l m )2
m m
ρm li lj
Aij = , (165)
m
lm
and 0 is the volume of the undeformed crystal. In particular, for a cubic crystal,
the three independent elastic constants are given by
1
C11 = [B11 + E (ρe )W11 + E (ρe )A211 ], (166)
0
1
C12 = [B12 + E (ρe )W12 + E (ρe )A211 ], (167)
0
1
C44 = [B12 + E (ρe )W12 ]. (168)
0
These equations nicely demonstrate the effect of the interplay between the pair
potential Uij and the embedding energy E e . Clearly, if we remove Uij = U , we
obtain C11 = C12 and C44 = 0, which, for real solid materials, are wrong. On
the other hand, if we remove the embedding energy E e = E, we obtain the well-
known Cauchy relation, C12 = C44 , which does not hold for all materials, but only
for a certain class of them.
To use the EAM one must have the embedding energy, the pair potential, and the
atomic densities. In their original work, Daw and Baskes (1984) and Foiles et al.
(1986) used semi-empirical correlations for evaluating these quantities. To develop
such correlations, one takes advantage of known properties of these quantities. For
example, the embedding energy, when defined relative to the free-atom energy,
must vanish for vanishing electron density, and must have negative slope and
positive curvature (second derivative) for the background electron densities found
in typical metals. On the other hand, the pair-interaction term in Eq. (159) is purely
repulsive, and its origin is Coulombic. These observations lead to the following
equation
Zi (r)Zj (r)
Uij (r) = , (169)
r
524 9. Atomistic Modeling of Materials
where Zi (r) is the effective charge of atom i. Note that Zi (r) must be positive
and decrease monotonically with increasing r. A particularly simple, yet accurate,
expression is given by
Z(r) = Zo (1 + a1 r a2 ) exp(−a3 r), (170)
where Zo is the number of the outer electrons in the atom (for example, Zo = 10
for Ni, Pd, and Pt, and Zo = 11 for Cu, Ag, and Au). The parameters a1 , a2 and
a3 must be determined empirically, although a2 = 1 is accurate for Ni, Pd, and
Pt, while a2 = 2 leads to accurate elastic constants for Cu, Ag, and Au. Using
these observations and properties, Daw and Baskes (1984) and Foiles et al. (1986)
obtained very accurate empirical correlations for the embedding energy and the
effective charges for a variety of metals, from which they computed their various
properties, such as their elastic constants and surface energy; see also Johnson
(1988) who utilized the EAM to study FCC metals.
Holian et al. (1991) developed the following EAM for use in MD simulation of
deformation of materials under high stresses. The local embedding density ρi was
assumed to be given by a pairwise sum over all neighboring particles, weighted
by a spherical localization function w(rij ), such that
ρi = w(rij ), (171)
j =j
where
$ %2
1 rc2 − r 2
w(r) = , (172)
ed(d + 1) rc2 − re2
where d is the dimensionality of the system, e = exp(1), rc is the cutoff distance
given by Eq. (34), and re is the equilibrium nearest-neighbor distance. The pairwise
interaction potential U (r) was taken to be that given by Eq. (33), except that the
LJ part of the potential was written as
σ 12 σ 6
ULJ (r) = 4χ ε − , (173)
r r
where χ is the fractional pair-potential contribution to the total cohesive energy
(χ = 1/3 is a reasonable value for many metals). Finally, the embedding energy
Eie was taken to be
1
Eie (ρi ) = eεd(d + 1)(1 − χ )ρi ln ρi . (174)
2
One can also replace the LJ potential by the more flexible Morse-like potential,
UMorse = ε{exp[−α(r/re − 1)] − 1}2 − ε, (175)
where α, the steepness of the repulsive well, is related to 0 , the volume of the
undeformed system, and the bulk modulus of the system at equilibrium.
However, the atomic densities can also be computed by the Hartree–Fock ap-
proximation, and the embedding energy and the pair potentials can be calculated
by the ab initio method. The ab initio electronic structure calculations are utilized
9.7. Interatomic Interaction Potentials 525
where U2 , the two-body term, can include such effects as the steric repulsion,
charge transfer between atoms, charge-dipole and van der Waals interactions, and
therefore
Hij Z i Zj Dij Wij
U2 (rij ) = nij + exp(−rij /a) − 4 exp(−rij /b) − 6 . (177)
rij rij rij rij
The first term of Eq. (177) represents a two-parameter representation of the steric
repulsion; the second term is the Coulombic interaction due to charge transfer and
contains the effective atomic charges Zi and Zj ; the third term takes into account
the charge-dipole interaction due to large polarizability of negative ions, while the
last term corresponds to the induced dipole-dipole interactions. Covalent effects
are taken into account through three-body bond-bending and bond-stretching terms
(similar to the Keating model described in Chapter 8 of Volume I), and includes
the Si-C as well as C-Si-C bond angles. It is given by
$ %
γ γ (cos θj ik − cos θ̄j ik )2
U3 (rij , rik ) = Bij k exp − #(r0 − rij )#(r0 − rik ).
rij − r0 rik − r0 1 + C(cos θj ik − cos θ̄j ik )2
(178)
526 9. Atomistic Modeling of Materials
Here, Bj ik is the strength of the interaction, #(r0 − rij ) is the step function, θj ik
is the angle formed by rij and rik , and θ̄j ik is a constant. The constant C plays an
important role if the material undergoes structural transformation. In the original
Stillinger–Weber formulation, cos θ̄j ik was assumed to have a value of −1/3, but
more generally one can treat this term as an adjustable parameter. Note that the
“ideal” tetrahedral angle θ is such that cos θ = −1/3, so that the trigonometric part
of Eq. (178) clearly discriminates in favor of pairs of bonds emanating from i with
the desired geometry. It is clear that the Stillinger–Weber potential contains many
parameters which must be estimated by fitting the predictions obtained with the
potential to certain properties of the material. Shimojo et al. (2000) employed this
potential in their MD simulations of cubic SiC under isothermal-isobaric condi-
tions (see Section 9.2), using only 1728 atoms. Table 9.4 compares the predictions
of the MD simulations with the experimental data, and it is clear that the agreement
between the two sets is excellent. In addition, the volume-pressure relation was
computed for SiC in the zinc-blende structure, i.e., in a configuration with 4-fold
coordination. Figure 9.10 compares the MD simulation results with the experimen-
tal data, and it is clear that the agreement is again excellent, hence demonstrating
the significance and utility of an accurate interatomic potential: If the interatomic
potential is accurate, then MD simulations provide quantitative predictions for
materials’ properties.
9.7. Interatomic Interaction Potentials 527
The Stillinger–Weber potential is, by far, the most widely used potential. It
has been utilized in the study of clusters, lattice dynamics, bulk point defects, the
liquid and amorphous states, surface diffusion and reconstructions, Si(100) stepped
surfaces, the liquid-vapor and crystal-melt interfaces, pulsed melting of surfaces,
epitaxial growth from the vapor, liquid-phase epitaxy, and growth of amorphous
films via atom deposition, as well as calculation of mechanical properties that
was mentioned above. It has also been extended to Ge, sulfur, fluorine, and Si-F
materials.
Mistriotis et al. (1989) modified the Stillinger–Weber potential in order to cor-
rectly describe clusters with more than 6 atoms. The angular dependence of the
three-body term was modified, and a four-body term was also added.
where
Eij = Uc (rij )[aij Ur (rij ) + bij Ua (rij )]. (180)
Here Ei and Eij are, respectively, a site and a bond energy, the sums are over the
atoms of the system, and rij is the distance between atoms i and j . The function Ur
represents a repulsive pair potential, which includes the orthogonalization energy
when atomic wave functions overlap (see Section 9.4), while Ua is the attractive
pair potential associated with bonding. As already emphasized throughout this
chapter, in many applications short-ranged functions allow a tremendous reduction
in computational effort, and therefore a cutoff function Uc has been introduced to
limit the range of the potential. The function bij is a measure of the bond order,
and is assumed to be a monotonically decreasing function of the coordination of
atoms i and j . In addition, terms that act to limit the range of interaction to the
first neighbor shell are included in bij , and the function aij consists solely of such
range-limiting terms.
Ferrante et al. (1983) showed that a large number of calculated binding-energies
for solid cohesion and chemisorption can be mapped onto a single dimensionless
function using a three-parameter scaling, and Abell (1985) showed that this univer-
sal behavior can be well-explained by use of a Morse or Morse-like pair potential,
Eq. (175). Therefore, Tersoff (1988) proposed the following expressions for Ur
and Ua :
Ur (r) = A exp(−λ1 r), (181)
Ua (r) = −B exp(−λ2 r), (182)
whereas the cutoff function Uc was taken to be
⎧
⎨ 1, r <R−D
Uc (r) = 1
− 1
sin[π(r − R)/(2D)], R − d <r <R+D (183)
⎩ 2 2
0, r >R+D
The cutoff function (and its derivative) is continuous for all r, and varies between
0 and 1 in a small range around R, which is selected so as to include only the
first-neighbor shell for most structures. In effect, the potential has the form of a
Morse pair potential, ignoring the three-body and higher-order effects, but with a
bond-order parameter bij that depends on the local environment. The function bij
is given by
bij = (1 + βζijn )−1/2n , (184)
with
ζij = Uc (rik )g(θij k ) exp[λ33 (rij − rik )3 ], (185)
k=i,j
bij = bj i which, however, does not have any significant consequence. If, however,
one insists on a more symmetric form, the sum over pairs of atoms in Eq. (180)
can be replaced with a sum over bonds (i > j ), and then bij can be replaced
with a symmetric function, b̄ij = 12 (bij + bj i ). Tersoff (1988) also proposed the
following equation for the function aij ,
n −1/2n
aij = (1 + α n ηij ) , (187)
ηij = Uc (rik ) exp[λ33 (rij − rik )3 ], (188)
k=i,j
where α is typically small enough that aij 1, unless, of course, ηij is exponen-
tially large.
Subsequently, Tersoff (1989) modified his proposed potential in order to describe
multicomponent mixtures, and more specifically C-Si and Si-Ge mixtures. Silicon
carbide, in particular, has a wide range of applications, ranging from optoelectric
devices and engineering materials to the basic substrate for membranes that must
operate at high-temperatures (Suwanmethanond et al., 2000). The reason for its
popularity is that it has excellent chemical stability, good electronic properties,
and high stiffness and hardness. In Tersoff’s generalization, Eqs. (179) and (180)
remain the same, but the remaining expressions are modified to account for the
multicomponent nature of the system. Thus,
Ur (rij ) = Aij exp(−λij rij ), (189)
Ua (rij ) = −Bij exp(−µij rij ), (190)
⎧
⎨ 1, rij < Rij
Uc (rij ) = 1
+ 1
cos[π(rij − Rij )/(Sij − Rij )], Rij < rij < Sij (191)
⎩ 2 2
0, rij > Sij
where the various parameters are given by
bij = χij (1 + βini ζijni )−1/2ni , (192)
ζij = Uc (rik )ωik g(θij k ), (193)
k=i,j
2 −2
g(θij k ) = 1 + ci {di − [di2 + (hi − cos θij k )2 ]−1 }, (194)
1 1
λij = (λi + λj ), µij = (µi + µj ), (195)
2 2
! !
Aij = Ai Aj , Bij = Bi Bj , (196)
! !
Rij = Ri Rj , Sij = Si Sj . (197)
Note that the parameter χij strengthens or weakens the heteropolar bonds, and
therefore represents in some sense the chemistry of the mixture. Note also that
χii = 1 and χij = χj i , and that ωii = 1, although experience has indicated that
ωij = 1 is also a reasonable estimate. Compared with the case of a pure component,
530 9. Atomistic Modeling of Materials
the potential for mixtures is somewhat simpler, as the parameter D of Eq. (183)
has been eliminated.
All the parameters of these potentials have been estimated by fitting them to heat
of formation and the properties of the respective elements. For Si the parameter D
is about 0.15Å. The rest of the parameters are listed in Table 9.5. To provide the
reader with some sense of the accuracy of these potentials, we compare in Table
9.6 the predicted lattice constant and the three elastic constants of cubic SiC with
the experimental data. It is clear that, except for C44 , the agreement between the
theoretical predictions and the data is quite good. In addition, Mura et al. (1998)
studied the properties of Si1−x Cx compounds in the range 0.125 ≤ x < 0.875 by
ab initio QMD method of Car and Parrinello (1985) (see Section 9.4) and by MD
simulations using the Tersoff potentials, and found very good agreement between
the predictions of the two methods.
However, because of the large difference in bonding characteristics between
hydrogen and carbon (recall that H is monovalent whereas C has a valency of
up to 4), a set of parameters cannot be found that can adequately describe bond
energies for a large number of hydrocarbons. Furthermore, the Tersoff potentials
9.7. Interatomic Interaction Potentials 531
may play important roles. In order to obtain a better understanding of this com-
plex process, atomic-scale simulations of this phenomenon is of course desirable.
However, its QMD simulation is still too costly (in terms of the computation time),
and thus the classical MD computations with an accurate interatomic potential is
desirable. The potential developed by Brenner (1990) is intended for this pur-
pose. Its main use is for hydrocarbons, and it can take into account the effect of
intramolecular chemical bonding (which the Tersoff potentials are incapable of
doing) in many small hydrocarbon molecules as well as graphite and diamond
lattices.
In its spirit, Brenner’s formulation is similar to Tersoff’s. The binding energy
Eb for the hydrocarbon potential is given by
Eb = [Ur (rij ) − b̄ij Ua (rij )], (198)
i j >i
where, similar to the Tersoff potentials, Ua and Ur are the attractive and repulsive
part of the energy, and are given by
#
Dije Sij 2
Ua (rij ) = fij (rij ) exp − βij rij − Rij ,
e
(199)
Sij − 1 Sij
Dije Sij !
Ur (rij ) = fij (rij ) exp − 2Sij βij rij − Rije , (200)
Sij − 1
where Dije is the well depth, and Rije is the equilibrium distance, the value of which
is the same as re in the Morse potential, Eq. (175). In fact, for Sij = 2 the attractive
and repulsive potentials are essentially identical with the Morse potential. fij (rij )
are cutoff functions given by
⎧ (1)
⎪
⎨ 1, r ≤ Rij
(1) (2) (1) (1) (2)
fij (r) = 2 + 2 cos[π(r − Rij )/(Rij − Rij )], Rij < r < Rij
1 1
(201)
⎪
⎩ (2)
0, r ≥ Rij
It should be clear that the cutoff functions explicitly restrict the interactions to
nearest neighbors. The function b̄ij is given by
1 (t) (t) conj
b̄ij = (bij + bj i ) + Fij [Ni , Nj , Nij ], (202)
2
implying that, similar to the Tersoff potentials, b̄ij depends on the environment
around atoms i and j and implicitly contains many-body information. The function
Fij is a correction term which is used only for carbon-carbon bonds, and
⎧ ⎫−δi
⎨ ⎬
(H) (C)
bij = 1 + g(θij k )fik (rik ) exp αij k [(rij − Rije ) − (rik − Rik
e
)] + Hij [Ni , Ni ] .
⎩ ⎭
k =i,j
(203)
9.7. Interatomic Interaction Potentials 533
(C) (H)
Here Ni and Ni are the number of carbon and hydrogen atoms bonded to atom
(t) (C) (H) conj
i, Ni = Ni + Ni , Nij depends on whether a bond between carbon atoms
i and j is part of a conjugated system, and θij k is the angle between bonds ij and
ik. Similar to Fij , the function Hij is a correction term, and both functions are
used only for hydrocarbons.
The cutoff functions fij (r) are used for defining the various quantities. One has
(H)
Ni = fij (rij ), (204)
j ={H}
(C)
Ni = fij (rij ), (205)
j ={C}
(t)
where, for example, {C} denotes the set of the carbon atoms. Values of Ni for
neighbors of two carbon atoms involved in a bond are used for determining whether
the bond is part of a conjugated system. For example, if the neighbors are carbon
(t)
atoms that have a coordination number of less four (i.e., Ni < 4), the bond is
defined as part of a conjugated system. For a bond between carbon atoms i and j ,
conj
Nij = 1 + fik (rik )F (xik ) + fj l (rj l )F (xj l ), (206)
k( =i,j )={C} l(=i,j )={C}
with
⎧
⎪
⎨ 1, xik ≤ 2
F (xik ) = 1
+ 1
cos[π(xik − 2)], 2 < xik < 3 (207)
⎪
⎩
2 2
1, xik ≥ 3
and
(t)
xik = Nk − fik (rik ). (208)
conj
The function F (xik ) yields a continuous value of Nij as bonds form and break
conj
and as second-neighbor coordinations change. If Nij = 1, a bond is not part of a
conj
conjugated system and the function yields appropriate values, while for Nij ≥ 2
the bond is part of a conjugated system and parameters fitted to conjugated bonds
are used. Finally, to make the potential continuous, 2D and 3D cubic splines are
utilized for the functions Hij and Fij , respectively, for interpolating between values
at discrete numbers of neighbors. The function g(θ ) is very similar to that in the
Tersoff potentials, Eqs. (186) and (194). For example, for carbon one has
c02 c02
gC (θ) = a0 1 + 2 − 2 . (209)
d0 d0 + (1 + cos θ )2
It is clear that the Brenner potential has a large number of parameters that
must be fitted to experimental data. The procedure for doing this is to first fit
to systems consisting of only carbons and hydrogen. Parameters are then se-
lected for the mixed hydrocarbon system that produce additive bond energies.
534 9. Atomistic Modeling of Materials
Because the pair terms are first fitted to solid-state carbon structures, the equilib-
rium carbon-carbon distances and the stretching force constants for hydrocarbons
are completely determined by fitting to bond energies. To determine appropriate
energies for hydrocarbons with carbon-carbon bonds, additive bond energies for
single, double, conjugated double and triple carbon-carbon bonds, and carbon-
hydrogen bonds are determined from molecular atomization energies. Values of
the parameter δi of Eq. (203) for carbon and hydrogen turn out to be identical and
equal to 0.80469. Values of the other parameters are listed by Brenner (1990), a list
too long to be given here. The Brenner potential has been utilized successfully in
studying many phenomena, including reaction of hydrocarbon species on diamond
surfaces, mechanical properties of graphite sheet and nanotubes, and CVD of di-
amond films. Moreover, excellent agreement was found (Robertson et al., 1992)
between the predictions of MD simulations of energetics of nanoscale graphitic
tubules using the Brenner potential, and those of first-principle electronic structure
calculations using local-density functional described in Sections 9.1 and 9.4.
Despite its success, the Brenner potential does have certain limitations. For ex-
ample, it does not include the resonance effect of aromatics. Most importantly, the
long-range van der Waals and Coulombic interactions are not included explicitly
in the model, although such interactions play an important role in many materials.
Che et al. (1999) modified the Brenner potential to take such effects into account.
In their formulation, the total energy of the system is written as
E= UijV (rij ) + Pij (rij )UijNB (rij ) . (210)
i j >i
Here superscript V denotes the valence short-range terms (for example, those in the
Brenner potential), NB indicates the long-range non-bond part of the energy (for
example, the contribution of van der Waals or Coulombic forces), while Pij = Pj i
is a screening function that properly weights the NB contributions to the total
energy, and is given by
'
Pij = f UijV , UijV V
f Uik V
, Ukj , (211)
k=i,j
with
exp(−x 2 y 2 ), if x < 0 and y < 0
f (x, y) = (212)
1, otherwise.
As pointed out by Che et al. (1999), the screening function eliminates NB inter-
actions between two atoms i and j when they are directly bonded (i.e., UijV < 0)
or when they are both bonded to a common atom k (i.e., Uik V < 0 and U V < 0).
kj
In both cases the screening function Pij is negligibly small, and therefore the NB
interactions do not make improper contribution to the total energy. Using the re-
lation, Fαβ = −∂E/∂rαβ , it is now straightforward to show that the force Fαβ
9.7. Interatomic Interaction Potentials 535
i>j
(213)
− V V
Uik Ukj Pij UijVDW Uik
V V
Fkj,αβ + Ukj
V V
Fik,αβ ,
ij k
with
∂UijV NB
∂Uαβ
V
Fij,αβ =− , NB
Fαβ = . (214)
∂rαβ ∂rαβ
V represents the valence forces, such as bond stretching, bending,
In Eq. (213), Fαβ
and torsion, and is contributed by the Brenner-type potential. The second term
represents the NB forces between two properly screened atoms. The third and
fourth terms are due to forces arising from correlations between screened bonds
which, in most cases, are negligible. Therefore, if atoms i and j do not form a
valence bond and do not also form a bond with the same atom k, then usually
V
Fij,αβ = 0, and either UikV = 0 or U V = 0, leading to zero contribution. However,
kj
even if both atoms i and j bind to a common atom k or if they form a valence bond
directly, these terms will still make a negligible contribution to the total energy
because of the exponential screening factor, Eq. (212), since in this case either
V < 0 or U V < 0. Note that there is no restriction on this formulation. That is,
Uiα iβ
the specifics of the NB terms do not alter the general formulation of this potential.
Che et al. (1999) used this generalized interatomic potentials to study the energetics
and structures of a variety of materials, including graphite and molecular crystals
and bucky tubes, and Lim et al. (2003) utilized it for generating porous amorphous
carbon structure and investigating transport, chemisorption and separation of gases
in such materials.
of the emission of electrons and light that is observed in the vicinity of the
fracture tip (see Chapters 6 and 7), it is entirely possible that even the DFT
would fail as well. Therefore, only a detailed and patient comparison of theory
and experiment, which, as of the time of writing this book, has not yet been
performed, will be able to settle such doubts.
(2) As discussed in Chapters 6 and 7, in steady state, the energy consumed by
the fracture per unit length must equal the energy stored, per unit length to
the left. This statement, which remains true even for strains large enough that
the applicability of linear elasticity could be called into question, relies on
symmetry rather than the matched asymptotics of fracture mechanics, and
must therefore be reproduced by any MD simulation.
(3) The MD simulation must contain a complete description of the cohesive zone.
However, as more energy is fed to the fracture tip, as temperature rises, or if
one studies heterogeneous or ductile materials, the size of the cohesive zone
increases, and therefore the size of the system used in the MD simulation
must increase accordingly. One should not expect MD simulations to pro-
vide “easy” (i.e., without much effort, patience, and efficient computational
strategy) predictions for materials where the cohesive zone is of the order
of microns, let alone millimeters, as in the silicon sample described above.
In this regard, MD simulation of fracture of amorphous polymers on which
many experiments have focused (see Chapters 6 and 7) provide a particularly
great challenge.
In what follows we first discuss the early MD simulations that involved only
a few hundred or thousand of atoms. This will give the reader an idea about the
long road that has been travelled in order to arrive at the present state-of-the-
art of the MD simulation of dynamic fracture. We then describe and discuss the
recent advances and compare the results of large-scale MD simulation of dynamic
fracture with the relevant experimental data.
The first attempt for comparing the predictions of the continuum fracture me-
chanics with the results of MD simulations was made at the beginning of the 1980s.
Thomson et al. (1971) had presented evidence for lattice trapping, a phenomenon
in which a crack neither propagates nor heals, rather it remains stable until exter-
nal loads somewhat larger than the Griffith threshold (see Chapters 6 and 7) are
imposed on the system. The magnitude of the trapping range depends strongly on
the characteristics of atomic bonding of materials. Lattice trapping may also de-
pend on the direction in which the crack tip bonds are broken, and may therefore
be different for fracture propagation along different crystallographic directions.
Moreover, as described in Section 7.12, Rice and Thomson (1974) had developed
a criterion for the degree of brittleness of a material according to which a material
can be considered as brittle if a dislocation in the neighborhood of the fracture tip
cannot escape from the tip region. These predictions were put to test in the first
truly MD simulations that were carried out by Paskin et al. (1980,1981). We call
their computations “true” MD simulations because, unlike Ashurst and Hoover
(1976), they utilized the LJ potential for representing the interactions between the
atoms in a triangular lattice. In their simulation, a crack was inserted in the middle
of the lattice in order to initiate fracture propagation. The smallest crack in a MD
simulation is represented by a pair of atoms the bond between which has been cut,
so that the two atoms do not directly interact with each other. An external force
was then applied to the lattice, and Newton’s equations of motion were solved
in order to calculate the atomic positions, velocities, and forces. The cutoff rc
for the LJ potential was assumed to be slightly smaller than two lattice bonds at
equilibrium. Paskin et al. showed that the Griffith energy criterion (see Chapters
6 and 7) is incorrect for large cracks. Their MD simulations also indicated that
lattice trapping is a negligible effect, which they attributed to the long range of
the interaction potentials. However, in general one should expect lattice trapping
to disappear at temperatures much lower than room temperature, and therefore,
in order to observe this phenomenon, experiments and MD simulations must be
carried out at very low temperatures. The necessity of a low temperature explains
why no lattice trapping has yet been observed experimentally in either crystalline
or amorphous materials. Paskin et al.’s simulations also indicated that the Rice–
Thomson criterion for brittleness is valid at low temperatures (see below for more
discussion of the Rice–Thomson theory).
In addition to Paskin et al.’s simulations, interesting MD computations were
carried out by Soules and Busbey (1983) to study fracture of sodium silicate fiber
glass. Instead of interatomic forces that result from the LJ potentials, these authors
used the following semi-empirical equation for computing the interatomic forces,
$ %4
e 2 Zi Z j r
Fij (r) = 3.448aij exp(−3.448r) + 1− , (216)
r2 rc
where aij is a parameter of the model in erg, Zi is the charge of atom i, e is the
electron charge, and rc is the cutoff distance at which the interatomic forces vanish.
The exponential term of this equation represents a repulsive force, while the 1/r 2
9.8. Molecular Dynamics Simulation of Fracture Propagation 541
was observed, caused by dislocation emission from the tip of the fracture. This
result indicated the existence of an energy barrier for nucleation of the dislocation.
Cheung and Yip (1990) showed by detailed calculations that this energy barrier
could not be predicted by the continuum theory of Rice and Thomson (1974)
mentioned above. This issue was also investigated by Zhou et al. (1994) using
MD simulations, who proposed that the Rice–Thomson theory should be modified
to include the effect of tensile broken-bonds, if it is to correctly predict dislocation
emission.
Figure 9.12. Off-diagonal views of the cohesive zone during the time of the brittle-to-
ductile transition. Darker and lighter areas indicate, respectively, whether the speed of the
atoms is less than or greater than one twelfth of the longitudinal sound speed (after Abraham
et al., 1997b; courtesy of Dr. Farid F. Abraham).
Figure 9.13. The onset of fracture instability (left), in reduced time intervals of 7 and
beginning at 85. The right panels show the fracture zipzag, beginning at reduced time 220
(after Abraham et al., 1997a; courtesy of Dr. Farid F. Abraham).
develops when the crack velocity approaches 1/3 of the Rayleigh wave speed cR .
At higher crack velocities, the fracture either follows a wavy path or branches out,
with the anisotropy that is due to the large deformation at the tip of the propagating
fracture playing the dominant role in determining the path of the fracture. This is of
course in contrast with the conventional wisdom (see Chapter 7) which associates
the lowest energy surface as the favored cleavage direction. Figure 9.13 nicely
demonstrates these results. The simulations also produced dislocations emission
from the rapidly moving fracture tip after the onset of the crack growth rough-
ening, implying that the dislocations are the consequence rather than the cause
of the dynamic instability. The number of the dislocations emitted was dependent
upon the external loading. These results are all in excellent agreement with the
experimental observations described in Chapters 6 and 7.
Let us point out that fractures at zero temperature display a clear dynamic
instability at a critical energy flux. Up to this point, phonons are able to carry
away all excess energy. This instability does not take the form of a simple micro-
branching instability, partly because atomic bonds can easily rejoin above the main
crack line in a single component solid with no environmental impurities available.
Instabilities at room temperature have not yet been explored either numerically or
analytically.
ness exponent α depends on the speed of fracture propagation. At the initial stages
of fracture, when the crack tip propagated slowly, α 0.44. However, once the
speed of fracture propagation exceeded a certain limit, a crossover was observed
to a higher value, α 0.8. These results are in agreement with the experiments of
Bouchaud and Navéos (1995) which were described in Chapters 6 and 7. In another
effort by this group, Li et al. (1996) studied dynamic fracture in SiSe2 nanowires,
and found that fracture is initiated in an amorphous region of the surface of the
material, while multiple fractures start at the boundaries of the amorphous region.
Finally, Kalia et al. (1997) investigated dynamic fracture in nanophase Si3 N4 ,
showing that intercluster regions of the material are amorphous, deflecting fracture
and hence giving rise to local crack branching. This implies that nanophase Si3 N4
can resist fracture much better than crystalline Si3 N4 . The roughness exponent
was found to be, α 0.84, in agreement with the experimental data.
loading) the crack tip first reaches the Rayleigh wave speed cR and for some time
propagates with this speed, and then jumps to a higher constant speed with a value
of about 8.97, essentially the same as the longitudinal wave speed cl . Similar re-
sults were obtained (Gumbsch and Gao, 1999) by MD simulations of propagation
of dislocation of indentation. These results are in complete agreement with the
experimental observations described in Section 7.8.15.
Let us point out that, while effective potentials, such as the embedded-atom and
LJ potentials (with fitted parameters), may be adequate for representing metals,
they are poor representatives of non-metallic materials. In this case, one must use
a first-principle quantum mechanical description of the materials in order to calcu-
late the potentials (see Sections 9.4 and 9.5). This is a rigorous method since, unlike
the case of MD simulations with empirical or semi-empirical potentials, ab initio
quantum mechanical computations do not use any adjustable parameters. More-
over, as discussed in Sections 9.2 and 9.4, the local density approximation in its
plane-wave pseudo-potential formulation can be optimized, so that the computa-
tions will be highly efficient. Kaxiras and Duesbery (1993) presented the results of
such a study for silicon, and Spence et al. (1993) used ab initio QMD simulations to
investigate the dependence of lattice trapping energies on applied load for fractures
propagating in silicon. Pérez and Gumbsch (2000) studied the anisotropy of cleav-
age fracture in silicon and the effect of sample size on the process. This approach
has not, however, been utilized extensively, presumably because its computational
cost is much larger than MD simulations using the empirical or semi-empirical
potentials discussed above.
Summary
Ab initio quantum mechanical computations based on the density functional the-
ory in the local density approximation, together with plane-wave pseudo-potential
formulation, offer an efficient and rigorous method for computing materials prop-
erties. Quantum MD simulation method of Car and Parrinello did not change the
essentials of such computations, but offered an enormous increase in the efficiency
of the method, hence making much larger pieces of materials accessible to such
computations.
As a technique for studying materials at the atomic scale, molecular dynamics
simulation has been used for several decades. However, development of vector
computers and parallel machines, and hence vectorized and parallelized compu-
tational algorithms, together with derivation of accurate interatomic potentials,
have made MD simulations a powerful tool for studying materials at atomic scales
utilizing millions of atoms and molecules.
These two computational strategies have enabled us to investigate and accurately
predict various properties of materials. We believe that when one or both of these
methods are joined with the multiscale approach that will be described in the next
chapter, the possibilities for accurate and efficient optimal design of materials with
specific properties may be limitless.
10
Multiscale Modeling of Materials:
Joining Atomistic Models with
Continuum Mechanics
10.0 Introduction
Throughout this book we have emphasized that solid materials of industrial impor-
tance are highly heterogeneous, with the heterogeneities manifesting themselves
at several length scales, ranging from the smallest to macroscopic scales. For many
centuries, such materials were discovered, mined, and processed in a serendipi-
tous manner. However, characterization of atoms and the progress made in x-ray
diffraction during the early decades of the twentieth century provided the impetus
for the search for a theory of materials that can explain their properties in terms of
their atomic constituents. It was soon realized that developing such a theory was
not practical yet, because
(1) the disparity between the relevant length scales, i.e., those at the atomic and
macroscopic scales is huge;
(2) the existing knowledge of the principles of atomic cohesion and the basic
properties of materials was totally inadequate, and
(3) the required computational power for solving the problem was (and still
is) enormous, while the available computational power was grossly limited.
Therefore, the consensus at the time was that any theory of materials could
not have predictive power.
Later decades of the twentieth century witnessed development of many quali-
tative and semi-quantitative models of materials that could explain the principles
of atomic cohesion, and basic properties of such fundamental materials as met-
als and semiconductors. Some of these models, though relatively simple, were
surprisingly accurate and helped us make remarkable progress in understand-
ing materials’ properties. However, as discussed in Chapter 9, for most materials
of industrial importance, the interatomic interactions are complex enough to re-
quire very elaborate models. Such models must usually be accompanied by very
extensive computations.
In Volume I, as well as in the present Volume, we utilized continuum mechan-
ics and lattice models to describe modeling and simulation of the morphology,
550 10. Multiscale Modeling of Materials
with lengths and widths that are several centimeters each and a thickness of about
a millimeter. Such samples would contain about 1022 atoms. The duration of the
experiment is about 50 µs whereas, as discussed in Chapter 9, the largest atomistic
simulations that are currently feasible can follow 108 atoms for around 10−9 s.
Direct atomistic simulation of fracture of such a sample of silicon will therefore
require more than eighteen orders of magnitude increase in computational power
over what we currently have. Such a breakthrough in computational power will
not be achieved any time soon, and therefore the critical question is: How can
one make a comparison between the results of the atomistic computations and the
relevant experimental data, or even the predictions of the continuum theories?
The essential problem facing such simulations is one of length and time scales.
However, before answering the above question, we should first remind ourselves
that the goal in computer simulations, both at the continuum and atomistic scales,
should not be performing the largest simulation utilizing the largest possible sys-
tem, but constructing the smallest one that is capable of answering specific physical
questions. In the example of brittle fracture of silicon, as well as in many other
phenomena that occur in materials, many important features may profitably be
studied by atomistic simulations that are comparatively small, involving only tens
of thousands of atoms. Therefore, a fruitful strategy may be based on merging
atomistic and continuum simulations. Hence, in the example of fracture of silicon,
one may describe the vicinity of the crack tip by atoms, but utilize the continuum
elasticity everywhere else.
The combined methodology is called a multiscale approach to modeling of
materials. In political jargon, multiscale modeling of materials is a divide-and-
conquer modeling paradigm. As the first step, the entire range of material behavior
is divided into a hierarchy of length scales. Next, the relevant physical processes
that are irreducible and operate independently at a given scale—sometimes referred
to as unit processes—are identified. The unit processes at one scale represent
averages of unit processes operating at the immediately lower length scale, and
this relation defines a partial ordering of the processes.
Over the past decade, multiscale modeling of materials has evolved from some-
thing that was thought of as a distant dream to a research area with intensive
methodological development. If its development continues at the current rapid
pace, it will, in the relatively near future, develop into a practical tool for indus-
trial applications involving design of materials that possess specific properties.
The goal of this chapter is to describe the basic ideas and techniques for multiscale
modeling of materials. Because the range of problems and phenomena to which
multiscale methods are applicable is very broad, we restrict the discussions to
the main theme of this book, namely, modeling of materials and predicting their
properties, and describe the essentials of the multiscale modeling approach by dis-
cussing a few recent applications. We begin our discussion by briefly describing
two main classes of multiscale modeling approach developed so far, after which
we describe some of their recent applications. More extensive discussions and
overviews pertaining to micromechanics and multiscale modeling of materials are
given by, for example, Ortiz and Phillips (1999) and Phillips (2001).
552 10. Multiscale Modeling of Materials
epitaxial thin film. Atomistic simulations can overcome such limitations. Using
either an ab initio or semi-empirical description of the interatomic interactions
between the material’s atoms, energy minimization or MD simulations, of the type
that were described in Chapter 9, can be utilized for investigating the structure,
energetics, and dynamics of a system consisting of an epitaxial film on a substrate.
Using a multiscale approach, Zepeda–Ruiz et al. (1999) studied the energetics,
strain fields, and semi-coherent interface structures in a layer-by-layer semicon-
ductor heteroepitaxy, such as InAs on GaAs(110) and InAs on GaAs(111)A, as well
as the interfacial stability with respect to misfit dislocation formation and the mor-
phology of the surface of the film grown on the substrate. The continuum theory
provided a parameterization scheme for the atomistic simulations. A Keating-type
potential (see Chapter 9), which contains the contributions of the stretching and
bond-bending forces, was utilized for representing the interatomic interactions,
and total energy minimization was used for determining the most stable configu-
ration of the system (i.e., the one with the lowest-energy state). The minimization
was done based on the conjugate-gradient method (see Section 9.5.2) with respect
to the atomic coordinates for a given strain state, which were uniaxial, biaxial, or
fully relaxed. Because of misfit dislocations a supercell (see Section 9.1.3) was
used that, depending on the state of the system, contained anywhere from 500 to
140,000 atoms. A major conclusion of this work was that the continuum theory of
elasticity can be accurate all the way down to the monolayer thickness, which is
the finest possible length scale for the theory in the context of layer-by-layer expi-
taxial growth. In addition, the theory was shown, in conjunction with the atomistic
simulations, to provide quantitative predictions for various properties of interest.
Indeed, even the linear isotropic elasticity was capable of fitting the results of the
atomistic simulations.
such computations alone cannot do justice to the problem, nor can they give us a
quantitative picture of what is happening during the deformation of a crystal, if
we consider the fact that macroscopic deformation of crystals involves dislocation
densities as high as 1013 m−2 , so that an area as small as 1 mm2 near the tip of a
crack can be crossed by over one hundred dislocations. In contrast, the supercell
that was used in the ab initio computations of Arias and Joannopoulos (1994)
contained only 324 atoms. Thus, it would be impossible to consider individual
dislocations at the macroscopic scale. Even the intermediate length scale, of the
order of a few hundred nanometres, is presently hardly within reach of conventional
atomistic simulation, as the number of atoms involved in such simulations would
be in excess of 108 , which makes purely atomistic simulations difficult to carry
out. One may develop phenomenological continuum models in which the defects
are treated as continuously distributed objects. However, such an approach could
not provide deep insight into the role of the defects in the properties of materials.
Tadmore et al. (1996a) were interested in studying the deformation processes
of interest at the intermediate length scale which, on one hand, involve discrete
dislocations in numbers that are too small to be described adequately by a con-
ventional continuum model of crystal plasticity, and, on the other hand, contain
too many atoms to be treatable by purely atomistic simulations. Thus, a multiscale
approach that joins the two methods, and would seem to be the only practical
solution, should have several key attributes, some of which are as follows.
(1) The theory should, at macroscopic length scales, reduce to the continuum
crystal elasticity and reproduce its important properties, such as material frame
indifference and crystal symmetry (see, for example, Milstein, 1982).
To achieve this goal, the atoms should be constrained to move in accordance
with the continuum displacement field. This would enable one to compute en-
ergies and forces from local lattice calculations. Thus, by construction, the
resulting continuum would automatically satisfy material frame indifference
and exhibit all the symmetries of the crystal. The system would also possess
lattice invariance, i.e., its energy would be invariant with respect to distor-
tions of the reference configuration. In particular, the energy density would
be periodic under crystallographic slip. An important consequence of this
periodicity is the lack of quasi-convexity of the energy functional [see, for
example, Fonseca (1988); see also Chapters 2 and 4, as well as Chapters 4,
7, and 10 of Volume I], which would make stable development of lattice de-
fects possible. The relaxation of functionals lacking quasi-convexity would
require consideration of minimizing sequences of deformations which exhibit
structure on increasingly finer scales, hence necessitating multiscale analysis
for simultaneous resolution of macroscopic and microscopic features into the
model.
(2) At the atomic scale, the theory should be built upon reliable interatomic inter-
actions, incorporate a lattice constant (or lattice parameter, as referred to by
Tadmor et al.), and possess all the invariance properties that are expected of a
crystal lattice.
10.2. Defects in Solids: Joining Finite-Element and Atomistic Computations 557
ment fields are constructed in the FE method. The selection of the representative
atoms may be based on the local variation of the deformation field. For example,
one may adapt the mesh in such a way that the variation of the displacement field
over each element of the triangulation does not exceed a fraction of the Burg-
ers vectors, hence ensuring that full atomistic resolution is achieved, for example,
near dislocation cores and on planes undergoing crystallographic slip. By contrast,
far away from any highly-stressed region, the density of the representative atoms
decreases rapidly, and the collective motion of a very large number of atoms is
governed, without significant loss of accuracy, by a small number of degrees of
freedom. In these coarse regions, the behavior of the model is ostensibly indistin-
guishable from that of a continuum. The effective equilibrium equations are then
obtained by minimizing the potential energy of the crystal over the reduced con-
figuration space. Therefore, the number of equilibrium equations that are obtained
is commensurate with the number of representative atoms. However, a direct cal-
culation of the effective force field requires, in principle, the evaluation of sums
that are extended over the full collection of atoms. Such full sums may be avoided
by the introduction of approximate summation rules, whereupon the complexity
of the computation of the effective force field becomes of the order of the reduced
model.
We now describe the details of the model. In the quasi-continuum (QC) theory
of Tadmor et al. (1996a) the continuum framework and continuum particle concept
are retained, but the macroscopic constitutive law is replaced by one based upon
direct atomistic calculations. The continuum particle is represented by a small
crystallite of radius Rc which surrounds a representative atom. This crystallite is
deformed according to the local continuum displacement field, and its energy is
computed based on an appropriate atomistic model. In order to compute the ener-
gies, one must make a correspondence between the deformation of the crystallite
and the continuum displacement field. A standard approach for doing so is based
on the Cauchy–Born rule (see, for example, Ericksen, 1984) according to which
the atomic positions are related to the continuum fields through a local deforma-
tion gradient F (which, for infinitesimal deformation, is the usual ∇u, where u is
the infinitesimal displacement) which is applied to the crystal’s undeformed lat-
tice basis. The crystal is then reconstructed from the altered base vectors. In this
manner each continuum particle is represented by an infinite crystal undergoing
homogeneous deformation. This limit is referred to as the local QC formulation.
The key idea is that, since the energy of each point is obtained directly from atom-
istic simulations, important properties of the crystal, such as its symmetries, are
automatically introduced into the description of the material. However, despite
being elegant and straightforward to implement, the local QC formulation suffers
from several shortcomings that make it necessary to develop a formulation that can
deal with non-local effects. Some of such shortcomings, as discussed by Tadmor
et al., are as follows.
(1) The most important shortcoming of the QC formulation is that, due to the
homogeneous nature of the deformation in this formulation, it is not possible
10.2. Defects in Solids: Joining Finite-Element and Atomistic Computations 559
Rc
Representative Atom
Figure 10.1. Local quasi-continuum/finite-element in which the triangle corners represent
the nodes, while the circles are atoms that belong to the crystallite (after Tadmor et al.,
1996a).
560 10. Multiscale Modeling of Materials
Rc
Representative Atom
Figure 10.2. Non-local quasi-continuum/finite-element (solid triangle) surrounded by
nearby elements (dashed triangles) (after Tadmor et al., 1996a).
Figure 10.3. Interfacial effects represented by a grain boundary (after Tadmor et al., 1996a).
local formulation also allows treatment of problems that involve, for example,
grain boundaries. If elements smaller than the representative crystallite radius Rc
are placed near such an interface, they will, due to non-locality, contain atoms that
are arranged in a different crystal orientation, and thus mimic a grain boundary;
see Figure 10.3.
10.2. Defects in Solids: Joining Finite-Element and Atomistic Computations 561
N
Fh (X, t) = a (t)∇ 0 Na (X), (6)
a=1
where a = 1, · · · , N denotes the nodes in the mesh, with N being the number of
nodes, a (t) are the nodal coordinates at time t, and Na (X) are the interpolation
functions. The main unknowns of the problem are now the nodal coordinates a (t)
which are obtained from the constrained minimization problem (4) in which the
ψ are replaced by ψ h , with the trial functions ψ h being of the following form,
N
ψ h (X) = ψ a Na (X). (7)
a=1
These trial functions must satisfy the boundary conditions identically on ∂B01 . All
integrals in Eq. (4), when written in terms of ψ h , can be conveniently computed
by numerical quadrature at the element level, which reduces all stress-strain calcu-
lations to the quadrature points of the elements. Tadmor et al. (1996a) used linear
three-noded triangular elements with one-point quadrature rule, and constructed
all the FE meshes by automatic triangulation based on the Delauney algorithm (see,
for example, Sloan, 1987). The constrained minimization problem was solved by a
conjugate-gradient approach (see Section 9.5.2), followed by a Newton–Raphson
iteration when the initial guess was too far from the solution.
within the cutoff radius rc . To account for this effect, Tadmor et al. introduced
the concept of an influence radius Ri associated with deformation F, which is
the radius that corresponds to the most distant point in the undeformed configura-
tion that is mapped onto the representative atom’s cut-off sphere in the deformed
configuration. Tadmor et al. showed that Ri is given by
!
Ri = rc λmax , (10)
where λmax is the largest eigenvalue of (F−1 )T F−1 . Then, every trial deformation
during the minimization process must satisfy the constraint, Ri ≤ Rc .
Given an acceptable trial deformation, the strain energy density Es , which is a
function of F, is computed using the atomistic method. The local contributions to
the out-of-balance force residual and global stiffness matrix follow from Eq. (4)
(when written in terms of ψ h ) as
∂E
local
= (P · ∇ 0 Na )d0 − ρ0 BNa d0 − T̄Na dS0 , (11)
∂ψ a e eh B0 ∂B02
∂ 2E
local
= [C : (∇ 0 Na ⊗ ∇ 0 Nb )]d0 , (12)
∂ψ a ∂ψ b e eh
where a and b are node numbers, P = ∂W/∂F is the first Piola–Kirchhoff stress
tensor (see Section 7.11.3), and C = ∂ 2 W/∂F2 is the Lagrangian tangent stiffness
tensor. These tensors are the finite deformation analogues of the usual Cauchy
stress and elastic modulus tensors in linear elasticity described in Chapter 7 of
Volume I. Note that the first terms of Eqs. (11) and (12) are sums only over local
elements e in B0 . In terms of the components of P and C, Eqs. (11) and (12) are
rewritten as
∂E
local
= PiJ Na,J d0 − ρ0 Bi Na d0 − T̄i Na dS0 , (13)
∂ψai e eh B0 ∂B02
∂ 2E
local
= [CiJ KL Na,J Nb,L ]d0 , (14)
∂ψai ∂ψbk e eh
where
∂Es −1
PiJ = = σij FJj , (15)
∂FiJ
∂ 2 Es −1 −1
CiJ KL = = (cij kl + δik σj l )FJj FLl , (16)
∂FiJ ∂FkL
where the relations between P and C on one hand, and their spatial counterparts,
the Kirchhoff stress σ and the spatial moduli c, on the other hand have been used,
with δik being the Kronecker delta. Tadmor et al. utilized a three-noded linear
element, which means that F was constant within each element. Thus, the integrals
in Eqs. (13) and (14) are computed by evaluating the integrands for the value of F
within each element and multiplying the result by the area of that element.
10.2. Defects in Solids: Joining Finite-Element and Atomistic Computations 565
The treatment of the problem presented so far is completely general and inde-
pendent of the atomistic model. The appropriate expressions for the components of
the stiffness tensor C were already given in Chapter 9, Eqs. (9.162)–(9.164). The
corresponding components of the spatial moduli tensor c are then computed from
Eq. (16). For a purely local formulation, the expressions for C and c represent a
complete constitutive description of the problem. However, as discussed above, a
purely local formulation is unable to capture non-uniform effects, such as stacking
faults. Therefore, one must resort to a non-local formulation which is described in
the next section. Note, however, that for a pure infinite crystal which is deformed
homogeneously, translational invariance reduces the general expression for the
EAM to that of a single atom and all of its neighbors that are within a pre-specified
cut-off radius rc .
where we have used a slightly different notation than what was used in Eq. (9.159)
in order to avoid confusion. As before, ρi is the local embedding density around
atom i. Then, the non-local contribution to the out-of-balance force residual is
given by
$ %
∂E
non−local
1 1 m ∂rem
m
= U (ρe )ρ (re ) + E (re ) eh , (20)
∂ψai e
m
2 ∂ψ i
a
566 10. Multiscale Modeling of Materials
where the sum over the elements e is over only non-local elements in B0 , eh is
the area of element e, and the total electron density ρe at the representative atom
of element e is given by [see Eq. (9.160)]
ρe = ρ(rem ). (21)
m
(25)
Similar to the total out-of-balance force residuals, FE meshes that contain both
local and non-local elements yield a stiffness matrix which is obtained by super-
position of both Eqs. (14) and (25). The inclusion of non-local elements near highly
deformed regions, such as stacking faults, completes the QC formulation.
(1) One computes the Lagrangian strain tensor for all elements in the FE mesh.
(2) Condition (28) is checked for all elements smaller than Rc in order to identify
non-locally strained elements.
(3) The energy of elements identified in (2) is computed by using the non-local
formulation of the problem. Only those elements that satisfy criterion (29) are
retained and considered as non-local.
(4) The elements that are non-local by proximity are located by identifying all
elements that are smaller than Rc and are within a distance Rc of an element
that satisfies both criteria (28) and (29).
(5) In addition, all surface and interfacial elements are computed non-locally.
Tadmor et al. (1996a) carried out this procedure at the start of each iteration
and integrated it into the solution process. As they pointed out, an important point
regarding this algorithm is that, to ensure convergence once an element is identified
as non-local, it must remain so from that point on even if in future iterations it no
longer satisfies the non-locality criterion, or is no longer in proximity to a non-local
element.
Figure 10.5. Finite-element mesh for initially distorted stacking faults (after Tadmor et al.,
1996a).
570 10. Multiscale Modeling of Materials
larger than Rc and are thus local. The elements of the second type are long and
narrow, and are considered non-local, but because their width is larger than Rc ,
they exhibit non-local effects only in the y-direction. Therefore, surface effects
play no role. The height of the narrow
√ elements was set equal to the interplanar
distance in the (111) direction, a0 / 3, and the global origin was selected such that
these elements fell between adjacent atomic planes; see Figure 10.6. Hence, the
elements that straddle the slip plane were allowed to act as a kinematical mecha-
nism for introducing slip. Thus, from an atomistic point of view, there is a jump
in displacement across the slip plane, as expected, while in the continuum there
exists a continuous linear variation in slip. However, because the energy is com-
puted atomistically, the manner in which the slip is distributed in the continuum
is unimportant. Five layers of the narrow elements were necessary to capture the
contributions of the four atomic planes adjacent to the slip plane to the SF energy;
see Figure 10.7. In this model, the SF energy is equal to the total energy per unit
thickness, divided by the width of the system. The model predicts that the unre-
laxed SF energy is identical to that obtained from the LS analysis. Following a
Newton–Raphson minimization, Tadmor et al. (1996a) found a relaxed SF energy
of about 6.2 meVÅ−2 which is smaller than, but comparable with, the LS value.
In the second analysis, the FE model was generated in the primary [1̄10]-[111]
coordinate system, and a deformation corresponding to a 1/2 Burgers vector slip
10.2. Defects in Solids: Joining Finite-Element and Atomistic Computations 571
was used. On relaxation, Tadmor et al. (1996a) found the SF energy to be the
same as in the previous Shockley partial analysis, and out-of-plane displacements
between 0.822 and 0.824 Å were observed in all nodes above the slip plane, which
are very close to the expected value.
#
"
[I10]
!
! "
[111]
FE FE
TB Regions
Processors N=1; É;N+M
Domain Decomposition
where the sum is over all occupied states No up to the Fermi level, while the second
sum is over all pairs of atoms of the repulsive potential Ur . The eigenvalues {}
correspond to the one electron states of a first-principles Hartree–Fock or density
functional calculation (see Sections 9.0, 9.1, and 9.4), and are obtained from a
576 10. Multiscale Modeling of Materials
where the matrix elements of H and S were calculated by reducing the equiva-
lent integrals within an extensive database of first-principles calculations to the
following parametric forms, expressed in terms of the pairwise functions hαβ and
sαβ ,
equations of motion, Broughton et al. (1999) used the Verlet algorithm (see Section
9.2.2) since it is easily augmented to handle multiple time scale MD simulations.
A time step of t = 5 × 10−16 s was used. The mass m of the silicon atom is
m = 4.6639 × 10−26 kg. One can accelerate evaluation of the SW energy and
its corresponding forces by taking advantage of atomic neighbor tables, so that
computer time scales as O(N ), where N is the number of atoms in the system.
The two- and three- body terms in the SW potential truncate smoothly to zero
just before the second-neighbor distance in zero-pressure diamond cubic structure
silicon.
where K and M are local stiffness and mass matrices, respectively, m is the cell
index, Nc is the total number of FE cells, u and u̇ defined only at the apices of each
triangle, and the sum over (p, q) is over the (3 × 2) Cartesian directions associated
578 10. Multiscale Modeling of Materials
with the same apices. The stiffness matrix K (m) associated with the mth triangle
is given by
L T (m)
K (m) = B CB , (40)
4Am
where Am is the area of the mth triangle, C is the reduced (3 × 3) elastic constant
matrix, B(m) is the matrix of coordinate differences of the apices of the FE mesh,
and L is the thickness of the material in the third dimension. C depends upon
the orientation of the system and is a function of the three basic elastic constants
of silicon, namely, C11 , C12 , and C44 , which are given for zero-temperature SW
silicon by Ray (1988) and Balamane et al. (1992) (these values are listed in Tables
9.4 and 9.6). Broughton et al. (1999) used the average of these quantities reported
in the literature. B(m) is given by
⎛ m ⎞
b1 0 a1m
⎜ ⎟
⎜ 0 a1m b1m ⎟
⎜ ⎟
⎜ bm 0 a m ⎟
⎜ 2 2 ⎟
[BT ](m) = ⎜ ⎟, (41)
⎜ 0 a2m b2m ⎟
⎜ ⎟
⎜ m ⎟
⎝ b3 0 a3m ⎠
0 a3m b3m
with
blm = yl+1
m
− yl+2
m
,
alm = xl+2
m
− xl+1
m
, (42)
where l = 1, 2 and 3 denotes the cyclic apex index, and x and y are the 2D FE
mesh coordinates with respect to which the displacements u were defined.
The mass matrix M requires some care since, in principle, the kinetic energy
density varies across any given cell. However, it is necessary to reduce the FE mesh
in the interface between FE and MD regions so as to coincide with the perfect
atomic lattice. Each atom is apportioned its kinetic energy accordingly. Thus,
for the FE mesh, Broughton et al. (1999) used the "lumped-mass" approximation
which reduces for the smallest mesh size to the atomic limit. In this approximation,
one third of the mass in each cell is apportioned to each apex, so that the kinetic
energy is given by
Nm
Ek = M t (u̇t )2 , (43)
t=1
ρL
Nc
3
Mt = δtml Am , (44)
3
m=1 l=1
where t labels the FE mesh points, of which there are Nm , and ml denotes the mesh
point index at each of the three apices of cell m. The u̇ are vectors of length two
since they relate to a mesh point.
10.3. Fracture Dynamics 579
Forces that correspond to the displacements in Eq. (39) were computed by taking
the spatial derivatives. Displacements and their time derivatives were obtained as
functions of time, for given boundary conditions, using the same update algorithm
and time step as those used for the MD and TB computations. The force due to the
mth cell is given by
(m)
fFE = K (m) u(m) , (45)
(m)
where fFE and u(m) are of length six. The total force associated with a mesh point
is then the sum of the contributions from each of the cells with apices in common
with that point, and
t
fFE = M t üt . (46)
To address the issue of the proper Hamiltonian HFE/MD , Broughton et al. (1999)
defined a conservative Hamiltonian so as to ensure symplectic time evolution (see
Chapter 9) of the atomic and displacement trajectories within the interface region.
To conceptualize this Hamiltonian, imagine that two different materials sit on
either side of an interface, such that on one side is FE silicon, whereas on the other
side one has SW silicon. The cross terms (i.e., the interface Hamiltonian HFE/MD )
can, to first order, be approximated by the average of the two descriptions: All FE
triangles that cross the interface contribute half their weight to the Hamiltonian,
while the triangles that are fully in the MD region contribute nothing. Similarly,
any SW interaction which crosses the interface contributes half its usual weight,
while the SW interaction between mesh points, which are fully on the FE side of the
interface, contributes nothing to HFE/MD . Since as discussed above, the atoms and
mesh points cannot be distinguished from each other, the SW energy formulation
for atomic coordinates r and the FE energy formulation for the displacements u
can be used throughout the interface. The one-to-one mapping of atoms to nodes is
no longer needed at distances (from the interface in the FE region) that are greater
than twice the SW pair cutoff, which is the distance of greatest three-body range
in the SW potential. Figure 10.12 presents such interactions and their ranges. Thus
1
Nx
6
I mI mI
EFE/MD = um
p Kpq uq
4 I
m =1 p,q=1
⎡ ⎤ (47)
1
+ ⎣ U2 (r(ij )I ) + U3 (r(ij )I , r(ik)I )⎦ ,
2 I (i<j ) [i,(j <k)]I
where U2 and U3 are the two- and three-body terms. Here, the superscript I refers
to those interactions that cross the FE/MD boundary [see Eq. (9.176)]. Indeed, as
discussed above, EFE/MD is defined only for interactions that cross the boundary.
Equation (47) allows any one atom of the triplet in the three-body terms to be on
an opposite side of the interface to the other two.
10.3. Fracture Dynamics 581
Since in the work of Broughton et al. (1999), the FE mesh was 2D while the MD
region was 3D, the third dimension of the FE region was treated by mean-field
approximation. Thus, in Eq. (47), x- and y-displacements of atoms on the MD
side of the FE/MD boundary that contribute to the elastic energy were obtained
by averaging over all equivalent atoms at the depth z. In a similar fashion, in
determining the SW energy contribution to HFE/MD , all x- and y-displacements in
the third dimension were replicated on the FE side of the boundary by assuming that
atoms are located at ideal lattice sites in that dimension. The overall Hamiltonian
remains conservative.
Two other issues regarding the definition of energy must still be addressed, both
of which involve a reference state. One involves the potential energy, while the
other has to do with the thermal energy. Consider first the potential energy. The
SW potential is referenced to infinitely separated atoms, whereas the FE potential
is referenced to a T = 0 unstrained lattice. Therefore, a constant offset energy that
did not affect the dynamics was added to each FE mesh point. The T = 0 energy
density for SW silicon at zero pressure is -4.33444 eV/atom. The offset energy was
computed for every FE point using an equation entirely analogous to that used to
compute mass in the “lumped-mass” approximation except that, instead of a mass
density, the SW energy density was used [see Eq. (43)]. This scheme ensured the
correct limiting behavior as the mesh spacing was reduced to atomic dimensions.
For atoms in the interface region, for systems with unusual orientation where
the offset is non-trivial to estimate atom by atom, a T = 0 calculation with zero
strain for the coupled FE/MD system can be performed. The offset may thereby
be calculated to maintain the energy/atom constant through the interface.
Rudd and Broughton (1998) showed that in the FE region (u̇)2 is related to the
temperature. However, as one moves away from the interface region and coarsens
the FE grid, atomic degrees of freedom are lost: The FE algorithm involves an
average over these degrees of freedom. Thus, to set the atomic and continuum
thermal energies on an equivalent level, the total FE thermal energy is written
again by using an offset. These corrected energies are denoted by a prime:
3 1
(E )FE = (Na − Nm )kB T + (Ek )FE + Nm kB T , (48)
2 2
3 1
(E )FE = (Na − Nm )kB T + EFE + Nm kB T . (49)
2 2
582 10. Multiscale Modeling of Materials
TB
TB TB
where the penultimate term involves only SW interactions crossing the boundary,
while the last term involves a TB calculation for the combined Si-UA system.
The above prescription produces a conservative Hamiltonian, if there is no
dynamic allocation of the TB region to those parts of the material where atomic
bonds are breaking. Unfortunately, for many materials and phenomena (such as
crack propagation) the 100 or so atoms, whose forces may currently be updated
using a non-orthogonal TB Hamiltonian in one second of real time, do not comprise
a large region. This problem may be partially addressed by using periodic boundary
conditions. One may also address the problem by using more than one processor
per TB region to perform the diagonalization, but such algorithms are presently
not efficient on coarse-grained scalable architecture computers. Broughton et al.
(1999) represented the region of breaking bonds by a region of TB segments, an
example of which is shown in Figure 10.14 for three overlapping TB regions. In
their simulations of fracture propagation (see below), Broughton et al. used eight
overlapping regions, with each region diagonalized separately, and also handled
by a separate processor. After forces on each atom are obtained for each TB region
separately, the force to be used in updating the velocity Verlet algorithm is obtained
as the average over the different regions; if there is no overlap of TB regions,
the same prescription as for a single TB region is used; where a UA of one TB
region overlaps a Si of another, the Si value is used. The number of atoms that are
propagated using TB forces is therefore less than the total number within all the
overlapping regions. These rules are, of course, intuitive.
In the simulation of fracture propagation, the energy and force algorithm, as
implemented in the MD and TB regions, proceeds by calculating the SW energy for
all atoms in the MD processors. The TB processors calculate not only TB energies
and forces, but also those SW forces that must be subtracted from those double
counted in MD processors. Thus, SW energies, by suitable division of two- and
three-body terms, are available for all atoms, which are then used to discriminate
different regions. The apex of a crack is found, for example, by locating the atom,
with a potential energy greater (more positive) than 60% of the bulk cohesive
potential energy, furthest to the left or right of the center of the system. The central
TB portion of the overlap region is then placed at that atom.
10.3. Fracture Dynamics 585
The slab was initialially at zero temperature, and a constant strain rate was
imposed on the outermost FE boundaries defining the opposing horizontal faces
of the slab. Moreover, a linear velocity gradient was applied within the slab, which
resulted in an increasing internal strain with time. The material failed at the notch
tip after it had been stretched by about 1.5%. The propagating cracks rapidly
achieved a limiting speed (2770 m/s) equal to 85% of the Rayleigh speed cR , the
sound speed of the solid Si surface. A very accurate signature of seamless coupling,
one that represents a validation of the method, is the fact that stress waves passed
from the MD region to the FE regions with no visible reflection at the FE/MD
interface, i.e., the coupling of the MD region to the FE region appeared seamless.
Moreover, there were no obvious discontinuities at the MD/TB interface.
the gas phase is regarded as a continuum, and phenomena at the micron length
scale, where the discrete nature of the gas phase becomes apparent and important.
For most CVD processes, flow and transport in the reactor at large are described
by the usual macroscopic conservation equations for momentum, mass, and energy
which give rise to coupled nonlinear partial differential equations that are typically
solved with the FE method. However, these continuum models are not valid as
length scales approach and shrink below the gas phase mean-free path λ. This is
an important consideration since microelectronic device features are frequently
submicron, whereas the mean-free path λ, under low pressure CVD conditions,
can be several hundred µm. This limitation on the use of continuum models is
widely recognized and has led to the development of discrete particle transport
models.
Rodgers and Jensen (1998) developed a multiscale model for CVD on length
scales ranging from microns to meters. At the macroscopic level the problems
of fluid flow and heat and mass transfer in a single wafer, low-pressure CVD
reactor were solved using the FE method. Since on the feature scale the continuum
models are no longer valid, transport at the feature scale was linked with the
continuum model by an effective reactivity function R that included effects of
both the multiscale surface heterogeneity and microscopic transport resistance.
A Monte Carlo method was then used for computing R for any set of reaction
pathways that occur over microelectronic device features of any geometry. Feature
scale computations were then combined to yield an effective reactivity map over
the surface of the substrate, which was subsequently utilized for formulating a flux
boundary condition for the continuum model. Iteration between macroscopic and
microscopic models was then used to ensure a consistent set of conditions at the
micro-macro interface.
For a different application of multiscale modeling methodology to problems
involving transport and reaction processes in nano- and microporous materials,
see Dadvar and Sahimi (2002, 2003).
Summary
The goal in multiscale modeling is to predict the performance and behavior of
heterogeneous materials in which there are several relevant and widely disparate
length scales. Due to the great complexities that are involved, the task of devel-
oping a multiscale model is usually referred to as a grand challenge problem. The
complexities that are involved are clear: At the atomic scale electrons govern the
interactions among atoms in a material, and thus quantum-mechanical effects are
important and must be taken into account. At the same time, at the macroscopic or
engineering scale, forces that arise from macroscopic stresses and/or temperature
gradients are the factors that control the performance of materials. At the interme-
diate length scales, defects such as dislocations control the mechanical properties
of materials up to tens of micrometers, while large collections of such defects,
including grain boundaries, govern their mesoscopic properties up to length scales
10.4. Other Applications of Multiscale Modeling 589
that are of the order of hundreds of micrometers. Coupling of all of these length
scales would not have been possible, had it not been for the tremendous advances in
the advent of ever more powerful, massively-parallel computers, and the great ad-
vances that have been made in the theoretical understanding of materials and their
properties. The emerging multiscale methodology demonstrates how coupling of
atomistic and continuum approaches results in more predictive power than either
approach offers alone.
An interesting and very useful aspect of multiscale modeling is the fact that, it
is a multidisciplinary field that brings together scientists from many disciplines.
The development of a multiscale model of any phenomenon that deals with one or
more aspect of materials’ properties should, in principle at least, involve chemists
and chemical engineers, applied physicists and mathematicians, and continuum
mechanicians.
References
Allen, M.P., and D.J. Tildesley, Computer Simulation of Liquids (Oxford University Press,
London, 1987).
Andersen, H.C., “Molecular dynamics simulations at constant pressure and/or temperature,”
J. Chem. Phys. 72, 2384 (1980).
Andersen, J.V., D. Sornette, and K.-t. Leung, “Tricritical behavior in rupture induced by
disorder,” Phys. Rev. Lett. 78, 2140 (1997).
Andrews, D.J., “Rupture velocity of plane strain shear cracks,” J. Geophys. Res. 81, 5679
(1976).
Anthony, J.B.C.S.R., J.B. Chubb, and J. Congleton, “The crack-branching velocity,” Philos.
Mag. 22, 1201 (1970).
Arakawa, K., and K. Takahashi, “Branching of a fast crack in polymers,” Int. J. Fract. 48,
245 (1991).
Aranson, I.S., V.A. Kalatsky, and V.M. Vinokur, “Continuum field description of crack
propagation,” Phys. Rev. Lett. 85, 118 (2000).
Arbabi, S., and M. Sahimi, “Elastic properties of three-dimensional percolation networks
with stretching and bond-bending forces,” Phys. Rev. B 38, 7173 (1988).
Arbabi, S., and M. Sahimi, “On three-dimensional elastic percolation networks with bond-
bending forces,” J. Phys. A 23, 2211 (1990a).
Arbabi, S., and M. Sahimi, “Test of universality for three-dimensional models of mechanical
breakdown in disordered solids,” Phys. Rev. B 41, 772 (1990b).
Archuleta, R.J., “Analysis og near-source static and dynamic measurements from the 1979
Imperial Valley earthquake,” Bull. Seismol. Soc. Am. 72, 1927 (1982).
Argon, A.S., “Brittle to ductile transition in cleavage fracture,” Acta Mettal. 35, 185 (1987).
Arias, T.A., and J.D. Joannopoulos, “Ab initio theory of dislocation interactions: From close-
range spontaneous annihilation to the long-range continuum limit,” Phys. Rev. Lett. 73,
680 (1994).
Ashcroft, N., and N.D. Mermin, Solid State Physics (Saunders, London, 1976).
Ashurst, W.T., and W.G. Hoover, “Microscopic fracture studies in the two-dimensional
triangular lattice,” Phys. Rev. B 14, 1465 (1976).
Åström, J.A., J.P. Mäkinen, M.J. Alva, and J. Timonen, “Elasticity of Poissonian fiber
networks,” Phys. Rev. E 61, 5550 (2000).
Åström, J.A., S. Saarinen, K.J. Niskanen, and J. Kurkijärvi, “Microscopic mechanics of
fiber networks,” J. Appl. Phys. 75, 2383 (1994).
Åström, J.A., and J. Timonen, “Crack bifurcations in a strained lattice,” Phys. Rev. B 54,
R9585 (1996).
Åström, J.A., and J. Timonen, “Fragmentation by crack branching,” Phys. Rev. Lett. 78,
3677 (1997a).
Åström, J.A., and J. Timonen, “Fracture of a brittle membrane,” Phys. Rev. Lett. 79, 3684
(1997b).
Auerbach, D.J., W. Paul, A.F. Bakker, C. Lutz, W.E. Rudge, and F.F. Abraham, “A special
purpose parallel computer for molecular dynamics: Motivation, design, implementation,
and application,” J. Phys. Chem. 91, 4881 (1987).
Baker, T.C., and F.W. Preston, “The effect of water on the strength of glasses,” J. Appl.
Phys. 17, 179 (1946).
Bakker, A.F., G.H. Gilmer, M.H. Grabow, and K. Thompson, “A special purpose computer
for molecular dynamics calculations,” J. Comput. Phys. 90, 313 (1990).
Balamane, H., T. Halicioglu and W. A. Tiller, “Comparative study of silicon empirical
interatomic potentials,” Phys. Rev. B 46, 2250 (1992).
References 593
Ball, R.C., and H. Larralde, "Three-dimensional stability analysis of planar straight cracks
propagating quasistatically under type I loading," Int. J. Fract. 71, 365 (1995).
Bao, G., J.W. Hutchinson, and R.M. McMeeking, “Particle reinforcement of ductile matrices
against plastic flow and creep,” Acta Metall. Mater. 39, 1871 (1991).
Barabási, A.-L., and H.E. Stanley, Fractal Concepts in Surface Growth (Cambridge
University Press, Cambridge, 1995).
Barabási, A.-L., and T. Vicsek, “Multifractality of self-affine surfaces,” Phys. Rev. A 44,
2730 (1990).
Baran, G.R., C. Roques-Carmes, D. Wehbi, and M. Degrange, “Fractal characteristics of
fracture surfaces,” J. Am. Ceram. Soc. 75, 2687 (1992).
Barber, M., J. Donley, and J.S. Langer, “Steady-state propagation of a crack in a viscoelastic
strip,” Phys. Rev. A 40, 366 (1989).
Barbosa, F.F., and S.L.A. de Queiroz, “Concentration anisotropy and directionality in the
dielectric breakdown problem on a square lattice,” J. Phys.: Condens. Matter 1, 2771
(1989).
Barclay, A.L., P.J.J. Sweeney, L.A. Dissado, and G.C. Stevens, “Stochastic modelling of
electrical treeing: fractal and statistical characteristics,” J. Phys. D 23, 1536 (1990).
Bardhan, K.K., “Nonlinear conduction in composites above percolation threshold-beyond
the backbone,” Physica A 241, 267 (1997).
Barenblatt, G.I., “The formation of equilibrium cracks during brittle fracture: general ideas
and hypothesis, axially symmetric cracks,” Appl. Math. Mech. (Translation of PMM) 23,
622 (1959a).
Barenblatt, G.I., “Concerning equilibrium cracks forming during battle fracture: the stability
of isolated cracks, relationship with energetic theories,” Appl. Math. Mech. (Translation
of PMM) 23, 1273 (1959b).
Barton, C.C., “Fractal analysis of the spatial clustering of fractures,” in Fractals and
Their Use in the Earth Sciences, edited by C.C. Barton and P.R. La Pointe (American
Geophysical Union, 1992), p. 126.
Barton, C.C., and P.A. Hsieh, Physical and Hydrological-Flow Properties of Fractures,
Guidebook T385 (American Geophysical Union, Las Vegas, Nevada, 1989).
Barton, C.C., and E. Larsen, “Fractal geometry of two-dimensional fracture networks at
Yucca Mountain, southwest Nevada,” in Proceedings of the International Symposium
on Fundamentals of Rock Joints, edited by O. Stephenson (Bjorkliden, Sweden, 1985),
p. 77.
Barton, C.C., T.A. Schutter, W.R. Page, and J.K. Samuel, “Computer generation of fracture
networks for hydrologic-flow modelling,” Trans. Amer. Geophys. Union 68, 1295 (1987).
Barton, J.L., “La relaxation diélectrique de quelques verres ternaires silice oxyde alcalin
oxyde alcalin-terreux,” Verres et Refr. 20, 328 (1966).
Baskes, M.I., J.S. Nelson, and A.F. Wright, “Semiempirical modified embedded-atom
potentials for silicon and germanium,” Phys. Rev. B 40, 6085 (1989).
Batrouni, G.G., and A. Hansen, “Fracture in three-dimensional fuse model,” Phys. Rev. Lett.
80, 325 (1998).
Bazant, M.Z., E. Kaxiras, and J.F. Justo, “Environment-dependent interatomic potential for
bulk silicon,” Phys. Rev B 56, 8542, (1997).
Beale, P.D., and P.M. Duxbury, “Theory of dielectric breakdown in metal-loaded
dielectrics,” Phys. Rev. B 37, 2785 (1988).
Beale, P.D., and D.J. Srolovitz, “Elastic fracture in random materials,” Phys. Rev. B 37,
5500 (1988).
594 References
Beazley, D.M., and P.S. Lomdahl, “Message-passing multi-cell molecular dynamics on the
Connection Machine CM-5,” Parallel Computing 20, 173 (1994).
Beazley, D.M., P.S. Lomdahl, N. Grφnbech-Jensen, R. Giles, and P. Tamayo, “Parallel
algorithms for short-range molecular dynamics,” Annual Rev. Comput. Phys. 3, 116
(1995).
Benguigui, L., “Simulation of dielectric failure by means of resistor-diode random lattices,”
Phys. Rev. B 38, 7211 (1988).
Benguigui, L., and P. Ron, “Laboratory simulation of dielectric breakdown,” in Non-
Linearity and Breakdown in Soft Condensed Matter, edited by K.K. Bardhan, B.K.
Chakrabarti, and A. Hansen, Lecture Notes in Physics 437 (Springer, Heidelberg, 1994),
p. 221.
Benguigui, L., P. Ron, and D.J. Bergman, “Strain and stress at the fracture of percolative
media,” J. Physique 48, 1547 (1987).
Bennett, C.H., “Efficient estimation of free energy differences from Monte Carlo data,” J.
Comput. Phys. 22, 245 (1976).
Beran, M.J., “Use of the variational approach to determine bounds for the effective
permittivity of random media,” Nuovo Cimento 38, 771 (1965).
Beran, M.J., Statistical Continuum Theories (Wiley, New York, 1968).
Bergkvist, H., “Some experiments on crack motion and arrest in polyethylmethacrylate,”
Eng. Fract. Mech. 6, 621 (1974).
Bergman, D.J., in Fragmentation, Form and Flow in Fractured Media, Ann. Israel Phys.
Soc. 8, 266 (1986).
Bergamn, D.J., “Nonlinear behavior and 1/f noise near a conductivity threshold: effects of
local microgeometry,” Phys. Rev. B 39, 4598 (1989).
Bergman, D.J., and Y. Kantor, “Critical properties of an elastic fractal,” Phys. Rev. Lett. 53,
511 (1984).
Bergman, D.J., O. Levy, and D. Stroud, “Theory of optical bistability in a weakly nonlinear
composite medium,” Phys. Rev. B 49, 129 (1994).
Bernstein, N., and E. Kaxiras, “Nonorthogonal tight-binding Hamiltonians for defects and
interfaces in silicon,” Phys. Rev. B 56, (1997).
Berry, M.V., “Regular and irregular semiclassical wavefunctions,” J. Phys. A 10, 2083
(1977).
Berveiller, M., and A. Zaoui, “An extension of the self-consistent scheme to plastically-
flowing polycrystals,” J. Mech. Phys. Solids 26, 325 (1979).
Bhattacharya, K., and R.V. Kohn, “Elastic energy minimization and recoverable strains of
polycrystalline shape-memory materials,” Arch. Rational Mech. Anal., (1997).
Bird, R.B., R.C. Armstrong, and O. Hasseger, Dynamics of Polymeric Liquids, 2nd ed.
(Wiley, New York, 1987).
Bishop, J.F.W., and R. Hill, “A theory of the plastic distortion of a polycrystalline under
combined stresses,” Philos. Mag. 42, 414 (1951a).
Bishop, J.F.W., and R. Hill, “A theoretical derivation of the plastic properties of a
polycrystalline face-center metal,” Philos. Mag. 42, 1298 (1951b).
Blumberg Selinger, R.L., Z.-G. Wang, W.M. Gelbart, and A. Ben-Shaul, “Statistical-
theormodynamic approach to fracture,” Phys. Rev. A 43, 4396 (1991a).
Blumberg Selinger, R.L., Z.-G. Wang, and W.M. Gelbart, “Effect of temperature and small-
scale defects on the strength of solids,” J. Chem. Phys. 95, 9128 (1991b).
Blumenfeld, R., and A. Aharony, “Nonlinear resistor fractal networks, topological distances,
singly connected bonds and fluctuations,” J. Phys. A 18, L443 (1985).
References 595
Coleman, B.D., “On the strength of classical fibres and fibre bundles,” J. Mech. Phys. Solids
7, 60 (1958).
Coppard, R.W., L.A. Dissado, S.M. Rowland, and R. Rakowski, “Dielectric breakdown in
metal-loaded polyethylene,” J. Phys.: Condens. Matter 1, 3041 (1989).
Cotterell, B., “Velocity effects in fracture propagation,” Appl. Mater. Res. 4, 227 (1965).
Cotterell, B., and A.G. Atkins, “A review of the J and I integrals and their implications for
crack growth resistance and toughness in ductile fracture,” Int. J. Fract. 81, 357 (1996).
Cotterell, B., and J.R. Rice, “Slightly curved or kinked cracks,” Int. J. Fract. 14, 155 (1980).
Cox, H.L., “The elasticity and strength of paper and other fibrous materials,” Br. J. Appl.
Phys. 3, 72 (1952).
Cracknell, R.F., D. Nicholson, and N. Quirke, “Direct molecular dynamics simulation of
flow down a chemical potential gradient in a slit-shaped micropore,” Phys. Rev. Lett. 74,
2463 (1995).
Crosby, K.M., and R.M. Bradley, “Fragmentation of thin films bonded to solid substrates:
simulation and a mean field theory,” Phys. Rev. E 55, 6084 (1997).
Curtin, W.A., “Theory of mechanical properties of ceramic-matrix composites,” J. Am.
Ceram. Soc. 74, 2837 (1991).
Curtin, W.A., “Toughening in disordered brittle materials,” Phys. Rev. B 55, 11270 (1997).
Curtin, W.A., M. Pamel, and H. Scher, “Time-dependent damage evolution and failure in
materials. II. Simulations,” Phys. Rev. B 55, 12051 (1997).
Curtin, W.A., and H. Scher, “Brittle fracture in disordered materials: A spring network
model,” J. Mater. Res. 5, 535 (1990a).
Curtin, W.A., and H. Scher, “Mechanics modeling using a spring network,” J. Mater. Res.
5, 554 (1990b).
Curtin, W.A., and H. Scher, “Analytic model for scaling of breakdown,” Phys. Rev. Lett.
67, 2457 (1991).
Curtin, W.A., and H. Scher, “Algebraic scaling of material strength,” Phys. Rev. B 45, 2620
(1992).
Curtin, W.A., and H. Scher, “Time-dependent damage evolution and failure in materials. I.
Theory,” Phys. Rev. B 55, 12038 (1997).
Dadvar, M., and M. Sahimi, “Pore network model of deactivation of immobilized glucose
isomerase in packed-bed reactors II: three-dimensional simulation at the particle level,”
Chem. Eng. Sci. 57, 939 (2002).
Dadvar, M., and M. Sahimi, “Pore network model of deactivation of immobilized glucose
isomerase in packed-bed reactors III: simulation at the reactor level,” Chem. Eng. Sci.
(2003).
Daguier, P., E. Bouchaud, and G. Lapasset, “Roughness of a crack front pinned by
microstructural obstacles,” Europhys. Lett. 31, 367 (1995).
Daguier, P., S. Henaux, E. Bouchaud, and F. Creuzet, “Quantitative analysis of a fracture
surface by atomic force microscopy,” Phys. Rev. E 53, 5637 (1996).
Daguier, P., B. Nghiem, E. Bouchaud, and F. Creuzet, “Pinning and depinning of crack
fronts in heterogeneous materials,” Phys. Rev. Lett. 78, 1062 (1997).
Dally, J.W., “Dynamic photoelastic studies of fracture,” Exp. Mech. 19, 349 (1979).
Dally, J.W., “Dynamic photoelasticity and its application to stress wave propagation,
fracture mechanics, and fracture control,” in Static and Dynamic Photoelasticity and
Caustics, edited by A. Lagarde (Springer, Berlin, 1987), p. 247.
Dally, J.W., W.L. Fourney, and G.R. Irwin, “On the uniqueness of the stress intensity factor-
crack velocity relationship,” Int. F. Fract. 27, 169 (1985).
References 599
Daniels, H.E., “The statistical theory of the strength of bundles of threads. I.,” Proc. R. Soc.
Lond. A 183, 404 (1945).
Dauskardt, R., F. Haubensak, and R.O. Ritchie, “On the interpretation of the fractal character
of fracture surfaces,” Acta Metall. Mater. 38, 143 (1990).
Daw, M.S., and M.I. Baskes, “Semiempirical, quantum mechanical calculation of hydrogen
embrittlement in metals,” Phys. Rev. Lett. 50, 1285 (1983).
Daw, M.S., and M.I. Baskes, “Embedded-atom method: Derivation and application to
impurities, surfaces, and other defects in metals,” Phys. Rev. B 29, 6443 (1984).
de Arcangelis, L., A. Hansen, H.J. Herrmann, and S. Roux, “Scaling laws in fracture,” Phys.
Rev. B 40, 877 (1989).
de Arcangelis, L., and H.J. Herrmann, “Scaling and multiscaling laws in random fuse
networks,” Phys. Rev. B 39, 2678 (1989).
de Arcangelis, L., S. Redner and H.J. Herrmann, “A random fuse model for breaking
processes,” J. Physique 46, L585 (1985).
deBotton, G., “The effective yield strength of fiber-reinforced composites,” lnter. J. Solids
Structures 32, 1743 (1995).
deBotton, G., and P. Ponte Castañeda, “On the ductility of laminated materials,” Inter. J.
Solids Structures 29, 2329 (1992).
deBotton, G., and P. Ponte Castañeda, “Elastoplastic constitutive relations for fiber-
reinforced solids,” Inter. J. Solids Structures 30, 1865 (1993).
deBotton, G., and P. Ponte Castañeda, “Variational estimates for the creep behavior of
polycrystals,” Proc. R. Soc. Lond. A 448, 121 (1995).
de Buhan, P., and A. Taliercio, “A homogenization approach to the yield strength of
composites,” Europ. J. Mech. A Solids 10, 129 (1991).
Dickinson, J.T., “Fracto-emission,” in Non-Destructive Testing of Fibre-Reinforced Plastics
Composites, edited by J. Summerscales, Vol. 2 (Elsevier, London, 1991).
Dienes, G.J., and A. Paskin, “Molecular dynamic simulations of crack propagation,” J. Phys.
Chem. Solids 48, 1015 (1987).
Ding, H.-Q., N. Karasawa, and W.A. Goddard III, “Atomic level simulations on a million
particles: the cell multipole method for Coulomb and London nonbond interactions,” J.
Chem. Phys. 97, 4309 (1992).
Dissado, L.A., and J.C. Fothergill, Electrical Degradation and Breakdown in Polymers
(Peregrinus, London, 1992).
Dissado, L.A., and P.J.J. Sweeney, “Physical model for breakdown structures in solid
dielectrics,” Phys. Rev. B 48, 16261 (1993).
Dobrodumov, A.M., and A.M. El’yashevich, “Simulation of brittle fracture of polymers by
a network model in the Monte Carlo method,” Sov. Phys.-Solid State 15, 1259 (1973).
Döll, W., “An experimental study of the heat generated in the plastic region of a running
crack in different polymenc materials,” Eng. Fract. Mech. 5, 259 (1973).
Döll, W., “A molecular weight dependent transition in polymethylmethacrylate,” J. Mater.
Sci. 10, 935 (1975).
Doyle, W.T., “The Clausius–Mossotti problem for cubic arrays of spheres,” J. Appl. Phys.
49, 795 (1978).
Dreizler, R.M., and E.K.U. Gross, Density-Functional Theory (Springer, Berlin, 1990).
Druker, D.C., “On minimum weight design and strength of non-homogeneous plastic bod-
ies,” in Non-homogeneity in Elasticity and Plasticity, edited by W. Olszag (Pergamon
Press, New York, 1959), p. 139.
Drucker, D.C., “The continuum theory of plasticity on the macroscale and the microscale,”
J. Mater. 1, 873 (1966).
600 References
Fisher, D.S., M.P.A. Fisher, and D.A. Huse, “Thermal fluctuations, quenched disorder, phase
transitions, and transport in type-II superconductors,” Phys. Rev. B 43, 130 (1991).
Fleck, N.A., and J.W. Hutchinson, “Strain gradient plasticity,” Adv. Appl. Mech. 33, 295
(1997).
Fleming, R.M., and C. C. Grimes, “Sliding-mode conductivity in NbSe3 : Observation of a
threshold electric field and conduction noise,” Phys. Rev. Lett. 42, 1423 (1979).
Flytzanis, C., Prog. Opt. 29, 2539 (1992).
Foiles, S.M., M.I. Baskes, and M.S. Daw, “Embedded-atom-method functions for fcc metals
Cu, Ag, Au, Ni, Pd, Pt, and their alloys,” Phys. Rev. B 33, 7983 (1986).
Fonesca, I., J. Math. Pure. Appl. 67, 175 (1988).
Form, W., N. Ito, and G.A. Kohring, “Vectorized and parallelized algorithms for multi-
million particle MD-simulation,” Int. J. Mod. Phys. C 4, 1075 (1993).
Fox, G.C., M.A. Johnson, G.A. Lyzenga, S.W. Otto, J.K. Salmon, and D.W. Walker, Solving
Problems on Cocurrent Processors, Vol. 1 (Prentice Hall, Englewood Cliffs, NJ, 1988).
Frenkel, D., and A.J.L. Ladd, “New Monte Carlo method to compute the free energy of
solids. Application to the FCC and HCP phases of hard spheres,” J. Chem. Phys. 81,
3188 (1984).
Freund, L.B., “The mechanics of dynamic shear crack propagation,” J. Geophys. Res. 84,
2199 (1979).
Freund, L.B., Dynamic Fracture Mechanics (Cambridge University Press, Cambridge,
1990).
Freund, L.B., and W.D. Nix, “A critical thickness condition for a strained compliant
substrate/epitaxial film system,” Appl. Phys. Lett. 69, 173 (1996).
Friel, J.J., and C.S. Pande, “A direct determination of fractal dimension of fracture surfaces
using scanning electron microscopy and stereoscopy,” J. Mater. Res. 8, 100 (1993).
Fuller, K.N.G., P.G. Fox, and J.E. Field, “The temperature rise at the tip of fast-moving
cracks in glassy polymers,” Proc. R. Soc. Lond. A 341, 1213 (1975).
Furukawa, H., “Propagation and pattern of crack in two dimensional dynamical lattices,”
Prog. Theor. Phys. 90, 949 (1993).
Gadenne, P., F. Brouers, V.M. Shalaev, and A.K. Sarychev, “Giant Stokes fields on
semicontinuous metal films,” J. Opt. Soc. Am. B 15, 68 (1998).
Galambos, J., The Asymptotic Theory of Extreme Order Statistics (Wiley, New York, 1978).
Gao, H., “Surface roughening and branching instabilities in dynamic fracture,” J. Mech.
Phys. Solids 41, 457 (1993).
Gao, L., B.-C. Xie, and Z.-Y. Li, “Crossover exponents in percolating nonlinear normal
conductor-insulator mixtures,” Physica A 271, 238 (1999).
Gărăjeu, M., and P. Suquet, “Effective properties of porous ideally plastic or viscoplastic
materials containing rigid particles,” J. Mech. Phys. Solids 45, 873 (1997).
Garboczi, E.J., “Linear dielectric-breakdown electrostatics,” Phys. Rev. B 38, 9005 (1988).
Gear, C.W., Numerical Initial Value Problems in Ordinary Differential Equations (Prentice-
Hall, Englewood Cliffs, NJ, 1971).
Gefen, Y., W.-H. Shih, R.B. Laibowitz, and J.M. Viggiano, “Nonlinear behavior near the
percolation metal-insulator transition,” Phys. Rev. Lett. 57, 3097 (1986).
George, A., and H. Michot, “Dislocation loops at crack tips: nucleation and growth—an
experimental study in silicon,” Mater. Sci. Eng. A 164, 118 (1993).
Germain, P., Q.S. Nguyen, and P. Suquet, “Continuum thermodynamics,” J. Appl. Mech.
50, 1010 (1983).
Gibiansky, L.V., and S. Torquato, “New approximation for the effective energy of non-linear
conducting composites,” J. Appl. Phys. 84, 301 (1998a).
References 603
Gibiansky, L.V., and S. Torquato, “Effective energy of nonlinear elastic and conduct-
ing composites: Approximations and cross-property bounds,” J. Appl. Phys. 84, 5969
(1998b).
Gilabert, A., S. Roux, and E. Guyon, “Current-voltage characteristic of a nonlinear resistor
network,” J. Physique 48, 1609 (1987).
Gilman, J.J., C. Knudsen, and W.P. Walsh, “Cleavage cracks and dislocations in LiF
crystals,” J. Appl. Phys. 6, 601 (1958).
Gilormini, “Insuffisance de l’extension classique du modéle auto-cohérent au comportement
non-linéaire,” C. R. Acad. Sci. Paris Ser. IIb 320, 115 (1995).
Goldstein, R.V., and R.K. Salganik, “Brittle fracture of solids with arbitrary cracks,” Int. J.
Fract. 10, 507 (1974).
Golosovsky, M., M. Tsindlekht, D. Davidov, and A.K. Sarychev, “Effective-medium ap-
proach to the microwave properties of the high-Tc superconductor-insulator composite,”
Physica C 209, 337 (1993).
Gordon, J.E., Structures or Why Things Don’t Fall Down (Penguin, London, 1978).
Gorkov, L.P., and G. Grüner (eds.), Charge Density Waves in Solids (Elsevier, Amsterdam,
1989).
Gouldstone, A., H.J. Koch, K.Y. Zeng, A.E. Ginnakopoulos, and S. Suresh, “Discrete and
continuous deformation during nanoindentation of thin films,” Acta Mater. 48, 2277
(2000).
Greengard, L., and V.I. Rokhlin, “A fast algorithm for particle simulations,” J. Comput.
Phys. 73, 325 (1987).
Grest, G.S., B. Dünweg, and K. Kremer, “Vectorized link cell Fortran code for molecular
dynamics simulations for a large number of particles,” Comput. Phys. Commun. 70, 243
(1989).
Griffith, A.A., “The phenomena of rupture and flow in solids,” Philos. Trans. R. Soc. Lond.
227, 163 (1920).
Gross, S., Dynamics of Fast Fracture, Ph.D. Thesis, University of Texas, Austin (1995).
Gross, S.P., J. Fineberg, M. Marder, W.D. McCormick, and H.L. Swinney, “Acoustic
emissions from rapidly moving cracks,” Phys. Rev. Lett. 71, 3162 (1993).
Gu, G.Q., and K.W. Yu, “Effective conductivity of nonlinear composites,” Phys. Rev. B 46,
4502 (1992).
Gumbel, E.J., Statistics of Extremes (Columbia University Press, New York, 1958).
Gumbsch, P., and H. Gao, “Dislocation faster than the speed of sound,” Science 283, 965
(1999).
Gumbsch, P., S.L. Zhou, and B.L. Holian, “Molecular dynamics investigation of dynamic
crack stability,” Phys. Rev. B 55, 3445 (1997).
Günes, P., S. Simsek, and S. Erkoc, “A comparative study of empirical potential energy
functions: Application to clusters,” Int. J. Mod. Phys. C 11, 451 (2000).
Gyure, M.F., and P.D. Beale, “Dielectric breakdown of a random array of conducting
cylinders,” Phys. Rev. B 40, 9433 (1989).
Gyure, M.F., and P.D. Beale, “Dielectric breakdown in continuous models of metal-loaded
dielectrics,” Phys. Rev. B 46, 3736 (1992).
Haile, J.M., and S. Gupta, “Extensions of molecular dynamics simulation method. II.
Isothermal systems,” J. Chem. Phys. 79, 3067 (1983).
Hamann, D.R., M. Schlüter, and C. Chiang, “Norm-conserving pseudopotentials,” Phys.
Rev. Lett. 43, 1494 (1979).
604 References
Hammonds, K.D., I.R. McDonald, and D.J. Tildesley, “Computational studies of the struc-
ture of carbon dioxide monolayers physisorbed on the basal plane of graphite,” Mol.
Phys. 70, 175 (1990).
Hansen, A., E.L. Hinrichsen, and S. Roux, “Roughness of crack interfaces,” Phys. Rev. Lett
66, 2476 (1991b).
Hansen, A., S. Roux, and E.L. Hinrichsen, “Annealed model for breakdown processes,”
Europhys. Lett. 13, 517 (1990).
Hansen, A., S. Roux, and H.J. Herrmann, “Rupture of central-force lattices,” J. Phys. France
50, 733 (1989).
Hansen, J.P., and McDonald, I.R., Theory of Simple Liquids (Academic Press, New York,
1986).
Harlow, D.G., Proc. R. Soc. Lond. A 397, 211 (1985).
Harlow, D.G., and S.L. Phoenix, “The chain-of-bundles probability model for the strength
of fibrous materials. I.Analysis and conjectures,” J. Comput. Mater. 12, 195 (1978).
Harlow, D.G., and S.L. Phoenix, “Approximations for the strength distribution and size
effect in an idealized lattice model of material breakdown,” J. Mech. Phys. Solids 39,
173 (1991).
Harris, A.B., “Field-theoretics formulation of the randomly diluted nonlinear resistor
network,” Phys. Rev. B 35, 5056 (1987).
Harris, A.B., T.C. Lubensky, W.K. Holcomb, and C. Dasgupta, “Renormalization-group
approach to percolation,” Phys. Rev. Lett. 35, 327 (1974).
Harris, R.A., and L.R. Pratt, “Discretized propagators, Hartree, Hartree–Fock equation, and
the Hohenberg–Kohn theorem,” J. Chem. Phys. 82, 856 (1985).
Hashin, Z., “The elastic moduli of heterogeneous materials,” J. Appl. Mech. 29, 143 (1962).
Hashin, Z. “Failure criteria for unidirectional fiber composites,” J. Appl. Mech. 47, 329
(1980).
Hashin, Z., and S. Shtrikman, “On some variational principles in anisotropic and
nonhomogeneous elasticity,” J. Mech. Phys. Solids 10, 335 (1962a).
Hashin, Z., and S. Shtrikman, “A variational approach to the theory of the elastic behavior
of polycrystals,” J. Mech. Phys. Solids 10, 343 (1962b).
Hashin, Z., and S. Shtrikman, “A variational approach to the theory of the elastic behavior
of multiphase materials,” J. Mech. Phys. Solids 11, 127 (1963).
Hassold, G.N., and D.J. Srolovitz, “Brittle fracture in materials with random defects,” Phys.
Rev. B 39, 9273 (1989).
Hauch, J., and M. Marder, “Energy balance in dynamic fracture, investigated by the potential
drop technique,” Int. J. Fract. (1999).
Heffelfinger, G.S., and F. van Swol, “Diffusion in Lennard–Jones fluids using dual control
volume grand canonical molecular dynamics simulation (DCV-GCMD),” J. Chem. Phys.
100, 7548 (1994).
Heiba, A.A., M. Sahimi, L.E. Scriven, and H.T. Davis, “Percolation theory of two-phase
flow in porous media,” Society of Petroleum Engineers paper 11015, New Orleans, LA
(1982).
Heiba, A.A., M. Sahimi, L. E. Scriven, and H. T. Davis, “Percolation theory of two-phase
relative permeability,” SPE Reservoir Engineering 7, 123 (1992).
Heino, P., and K. Kaski, “Mesoscopic model of crack branching,” Phys. Rev. B 54, 6150
(1996).
Heino, P., and K. Kaski, “Mesoscopic Maxwell-dissipative finite element model for crack
propagation,” Int. J. Mod. Phys. C 8, 383 (1997).
References 605
Heinrichs, J., and N. Kumar, “Simple exact treatment of conductance in a random Bethe
lattice,” J. Phys. C 8, L510 (1975).
Hellmann, H., Einführung in die Quantumchemie (Deuticke, Leipzig, 1937).
Hendrickson, B., and R. Leland, “An improved spectral graph partitioning algorithm for
mapping parallel computations,” SIAM J. Sci. Stat. Comput. 16, 452 (1995).
Herrmann, H.J., A. Hansen, and S. Roux, “Fracture of disordered, elastic lattices in two
dimensions,” Phys. Rev. B 39, 637 (1989a).
Herrmann, H.J., J. Kertész, and L. de Arcangelis, “Fractal shapes of deterministic cracks,”
Europhys. Lett. 10, 147 (1989b).
Herrmann, H.J., and S. Roux (eds.), Statistical Models for the Fracture of Disordered Media
(North-Holland, Amsterdam, 1990).
Herrmann, H.J., and M. Sahimi, “Fluid penetration through a crack in a pressure gradient,”
J. Phys. A 26, L1145 (1993).
Herrmann, H.J., M. Sahimi, and F. Tzschichholz, “Examples of fractals in soil mechanics,”
Fractals 1, 795 (1993).
Hervé, E., C. Stolz, and A. Zaoui, “A propos de l’assemblage des sphéres composites de
Hashin,” C. R. Acad. Sci. Paris Ser. II 313, 857 (1991).
Hesselbo, B., D. Phil. Thesis, Oxford University (1994).
Hill, R., “Elastic properties of reinforced solids: Some theoretical principles,” J. Mech.
Phys. Solids 11, 357 (1963).
Hill, R., “Continuum micro-mechanics of elastoplastic polycrystals,” J. Mech. Phys. Solids
13, 89 (1965a).
Hill, R., “Self-consistent mechanics of composite materials,” J. Mech. Phys. Solids 13, 213
(1965b).
Hill, R.M., and L.A. Dissado, “Theoretical basis for the statistics of dielectric breakdown,”
J. Phys. C 16, 2145 (1983a).
Hill, R.M., and L.A. Dissado, “Examination of the statistics of dielectric breakdown,” J.
Phys. C 16, 4447 (1983b).
Hinrichsen, E.L., A. Hansen, and S. Roux, “A fracture growth model,” Europhys. Lett. 8, 1
(1989).
Hirsch, P.B., R.W. Horne, and M.J. Whelan, “Direct observation of the arrangement and
motion of dislocation in aluminum,” Philos. Mag. (series 8) 1, 677 (1956).
Hirsch, P.B., J. Samuels, and S.G. Roberts, “The brittle-ductile transition in silicon. II.
Interpretation,” Proc. R. Soc. Lond. A 421, 25 (1989).
Ho, P.S., and T. Kwok, “Electromagnetism in metals,” Rep. Prog. Phys. 52, 301 (1989).
Hoagland, R.G., M.S. Daw, S.M. Foiles, and M.I. Baskes, “An atomic model of crack tip
deformation in aluminum using an embedded atom potential,” J. Mater. Res. 5, 313
(1990).
Hockney, R.W., S.P. Goel, and J.W. Eastwood, “Quite high-resolution computer models of
a plasma,” J. Comput. Phys. 14, 48 (1974).
Hodgdon, J.A., and J.P. Sethna, “Derivation of a general three-dimensional crack-
propagation law: A generalization of the principle of local symmetry,” Phys. Rev. B
47, 4831 (1993).
Hoeing, A., “Some implications of an elastic-electrostatic analogy on certain path-
independent integrals,” Int. J. Eng. Sci. 22, 87 (1984).
Hohenberg, P., and W. Kohn, “Inhomogeneous electron gas,” Phys. Rev. 136, B864 (1964).
Holian, B.L., and R. Ravelo, “Fracture simulations using large-scale molecular dynamics,”
Phys. Rev. B 51, 11275 (1995).
606 References
Holian, B.L., A.F. Voter, and R. Ravelo, “Thermostatted molecular dynamics: How to avoid
the Toda demon hidden in Nosé-Hoover dynamics,” Phys. Rev. E 52, 2338 (1995).
Holian, B.L., A.F. Voter, N.J. Wagner, R.J. Ravelo, S.P. Chen, W.G. Hoover, C.G. Hoover,
J.E. Hammerberg, and T.D. Dontje, “Effects of pairwise versus many-bondy forces on
high-stress plastic deformation,” Phys. Rev. A 43, 2655 (1991).
Holland, D., and M. Marder, “Ideal brittle fracture of silicon studied with molecular
dynamics,” Phys. Rev. Lett. 80, 746 (1998).
Hoover, W.G., “Canonical dynamics: Equilibrium phase-space distributions,” Phys. Rev. A
31, 1695 (1985).
Hoover, W.G., and F.H. Ree, “Melting transition and communal entropy for hard spheres,”
J. Chem. Phys. 49, 3609 (1968).
Horowitz, G.E., Arch. Elektrotech. Berlin 18, 555 (1927).
Hough, S.E., “On the use of spectral methods for the determination of fractal dimension,”
Geophys. Res. Lett. 16, 673 (1989).
Hua, L., H. Rafii-Tabar, and M. Cross, Philos. Mag. Lett. 75, 237 (1997).
Hughes, B.D., Random Walks and Random Environments, Vol. 1 (Oxford University Press,
London, 1995).
Hui, P.M., “Effective nonlinear response in dilute nonlinear granular materials,” J. Appl.
Phys. 68, 3009 (1990a).
Hui, P.M., “Enhancement in nonlinear effects in percolating nonlinear resistor networks,”
Phys. Rev. B 41, 1673 (1990b).
Hui, P.M., “Crossover electric field in percolating perfect-conductor-nonlinear-normal-
metal composites,” Phys. Rev. B 49, 15344 (1994).
Hui, P.M., P. Cheung, and Y.R. Kwong, “Effective response in nonlinear random
composites,” Physica A 241, 301 (1997).
Hull, D., “Effect of Crazes on the propagation of cracks in polystyrene,” J. Mater. Sci. 5,
357 (1970).
Hull, D., Fractography (Cambridge University Press, Cambridge, 1999).
Hull, D., and P. Beardmore, “Velocity of propagation of cleavage cracks in tungsten,” Int.
J. Fract. Mech. 2, 468 (1966).
Huntington, H.B., “Electromigration in Metals,” in Diffusion in Solids-Recent Develop-
ments, edited by A.S. Novick and J.J. Burton (Academic Press, New York, 1975),
p. 120.
Imre, A., T. Pajkossy, and L. Nyikos, “Electrochemical determination of the fractal
dimension of fractured surfaces,” Acta Metall. Mater. 40, 1819 (1992).
Inglis, C.E., “Stresses in a plate due to the presence of cracks and sharp corners,” Trans.
Inst. Naval Archit. 55, 219 (1913).
Irwin, G.R., “Analysis of stresses and strains near the end of a crack traversing a plate,” J.
Appl. Mech. 24, 361 (1956).
Irwin, G.R., “Fracture,” in Handbuch der Physik, Vol. 6 (Springer, Berlin, 1958), p. 551.
Irwin, G.R., J.W. Dally, T. Kobayashi, W.L. Fourney, M.J. Etheridge, and H.P. Rossmanith,
“On the determination of the a-K relationship for birefringent polymers,” Exp. Mech.
19, 121 (1979).
Jackson, J.D., Classical Electrodynamics, 3rd Ed. (John Wiley & Sons, 1998).
Janak, J.F., “Proof that ∂E/∂ni = i in density-functional theory,” Phys. Rev. B 18, 7165
(1978).
Jarvinen, R., T. Mantyla, and P. Kettunen, “Improved adhesion between a sputtered alumina
coating and a copper substrate,” Thin Solid Films 114, 311 (1984).
References 607
Jellinek, J., and R.S. Berry, “Generalization of Nosé’s isothermal molecular dynamics,”
Phys. Rev. A 38, 3069 (1988).
Joannopoulos, J.D., P. Bash, and A. Rappe, Chemical Design Automation News 6 (No. 8)
(1991).
Johnson, E., “Process region changes for rapidly propagating cracks,” Int. J. Fract. 55, 47
(1992).
Johnson, E., “Process region influence on energy release rate and crack tip velocity during
rapid crack propagation,” Int. J. Fract. 61, 183 (1993).
Johnson, H.T., and L.B. Freund, “Mechanics of coherent and dislocation island morpholo-
gies in strained epitaxial material systems,” J. Appl. Phys. 81, 6081 (1997).
Johnson, J.W., and D.G. Holloway, “On the shape and size of the fracture zones on glass
fracture surfaces,” Philos. Mag. 14, 731 (1966).
Johnson, J.W., and D.G. Holloway, “Microstructure of the mist zone on glass fracture
surfaces,” Philos. Mag. 17, 899 (1968).
Johnson, R.A., "Interstitials and vacancies in α-iron", Phys. Rev. 134A, 1329 (1964).
Johnson, R.A., “Analytic nearest-neighbor model for fcc metals,” Phys. Rev. B 37, 3924
(1988).
Jones, R.O., and O. Gunnarsson, “The density functional formalism, its applications and
prospects,” Rev. Mod. Phys. 61, 689 (1989).
Kahng, B., G.G. Batrouni, S. Redner, L. de Arcangelis, and H.J. Herrmann, “Electrical
breakdown in a fuse network with random, continuously distributed breaking strengths,”
Phys. Rev. B 37, 7625 (1988).
Kalia, R.K., A. Nakano, A. Omeltchenko, K. Tsuruta, and P. Vashishta, “Role of ultrafine
microstructures in dynamic fracture in nanophase silicon nitride,” Phys. Rev. Lett. 78,
2144 (1997).
Kallivayalil, J., and A.T. Zehnder, “A method for thermo-mechanical analysis of steady
state dynamic crack growth,” Int. J. Fract. 66 99 (1994).
Kalthoff, J.F., “On the propagation direction of bifurcated cracks,” in Dynamic Crack
Propagation, edited by G.C. Sih (Noordhoof, Leydon, 1972), p. 449.
Kalthoff, J.F., “The shadow optical method of caustics,” in Static and Dynamic
Photoelasticity and Caustics, edited by A. Lagarde (Springer, Berlin, 1987), p. 401.
Kane, E.O., “Phonon spectra of diamond and zinc-blende semiconductors,” Phys. Rev. B
31, 7865 (1985).
Kantor, Y., and I. Webman, “Elastic properties of random percolating systems,” Phys. Rev.
Lett. 52, 1891 (1984).
Kardar, M., “Fluctuations of interfaces and fronts,” in Disorder and Fracture, edited by
J.C. Charmet, S. Roux, and E. Guyon (Plenum, New York, 1990), p. 3.
Kardar, M., G. Parisi, and Y.C. Zhang, “Dynamic scaling of growing interfaces,” Phys. Rev.
Lett. 56, 889 (1986).
Karma, A., D.A. Kessler, and H. Levine, “Phase-field model of Mode III dynamic fracture,”
Phys. Rev. Lett. 87, 045501-1 (2001).
Kausch, H.H., Polymer Fracture (Springer, Berlin, 1987).
Kaveh, M., and N.F. Mott, “The conductivity of disordered systems and the scaling theory,”
J. Phys. C 14, L659 (1981).
Kawarabayashi, T., B. Kramer, and T. Ohtsuki, “Anderson transitions in three-dimensional
disordered systems with randomly varying magnetic flux,” Phys. Rev. B 57, 11842 (1998).
Kaxiras, E., and M.S. Duesbery, “Free energies of generalized stacking faults in Si and
implications for the brittle-ductile transition,” Phys. Rev. Lett. 70, 3752 (1993).
608 References
Kaxiras, E., and K.C. Pandey, “New classical potential for accurate simulation of atomic
processes in Si,” Phys. Rev. B 38, 12736 (1988).
Kelchner, C.J., S.J. Plimpton, and J.C. Hamilton, “Dislocation nucleation and defect
structure during surface indentation,” Phys. Rev. B 58, 11085 (1998).
Kellomäki, M., J. Åström, and J. Timonen, “Rigidity and dynamics of random spring
networks,” Phys. Rev. Lett. 77, 2730 (1996).
Kelly, A., W.R. Tyson, and A.H. Cotterell, “Ductile and brittle crystals,” Philos. Mag. 15,
567 (1967).
Kenkel, S.W., and J.P. Straley, “Percolation theory of nonlinear circuit elements,” Phys.
Rev. Lett. 49, 767 (1982).
Kent, R., and R. Rat, “Static electricity phenomena in the manufacture and handling of solid
propellants,” J. Electrostat. 17, 299 (1985).
Kerkhof, F., “Wave fractographic investigations of brittle fracture dynamics,” in Dynamic
Crack Propagation, edited by G.C. Sih (Noordhoff International Publishing, Leyden,
1973), p. 3.
Khanta, M., D.P. Pope, and V. Vitek, “Dislocation screening and the brittle-to-ductile
transition: A Kosterlitz-Thouless type instability,” Phys. Rev. Lett. 73, 684 (1994).
Kim, J.M., and J.M. Kosterlitz, “Growth in a restricted solid-on-solid model,” Phys. REv.
Lett. 62, 2289 (1989).
Kim, K.S., “Dynamic fracture under normal impact loading of the crack faces,” J. Appl.
Mech. 52, 585 (1985).
Kim, S., and S. Karrila, Microhydrodynamics. Principles and Selected Applications,
(Butterworth-Heinemann, Boston, 1991).
Kinloch, A.J., and R.J. Young, Fracture Behavior of Polymers (Applied Sciences, London,
1983).
Kinzel, W., “Directed percolation,” in Percolation Structures and Processes, edited by G.
Deutscher, R. Zallen, and J. Adler (Adam Hilger, Bristol, 1983), p. 425.
Knap, J., and M. Ortiz, “An analysis of the quasicontinuum method,” J. Mech. Phys. Solids
49, 1899 (2001).
Knaur, J.A., and P.P. Budenstein, IEEE Trans. EI-15, 313 (1980).
Kneipp, K., Y. Wang, H. Kneipp, L.T. Perelman, I. Itzkan, R. Dasari, and M.S. Feld, “Single
molecule detection using surface-enhanced Raman scattering (SERS),” Phys. Rev. Lett.
78, 1667 (1997).
Kobayashi, A., N. Ohtani, and T. Sato, “Phenomenological aspects of viscoelastic crack
propagation,” J. Appl. Poly. Sci. 18, 1625 (1974).
Kobayashi, A.S., in Handbook on Experimental Mechanics, edited by A.S. Kobayashi
(American Society for Testing and Materials, Prentice-Hall, Englewood Cliffs, NJ, 1987),
p. 231.
Kohlhoff, S., P. Gumbsch, and H. F. Fischmeister, “Crack propagation in BCC crystals
studied with a combined finite-element and atomistic model,” Philos. Mag. 64A, 851
(1991).
Kohn, R.V., and T.D. Little, “Some model problems of polycrystal plasticity with deficient
basic crystals,” (1997).
Kohn, R.V., and G.W. Milton, “On bounding the effective conductivity of anisotropic com-
posites,” in Homogenization and Effective Moduli of Materials and Media, edited by
J.L. Ericksen, D. Kinderlehrer, R.V. Kohn, and J.-L. Lions (Springer, New York, 1986),
p. 97.
Kohn, W., and L.J. Sham, “Self-consistent equations including exchange and correlation
effects,” Phys. Rev. 140, A1133 (1965).
References 609
Kohr, K.E., and S. Das Sarma, “Model-potential study of (2n + 1) × (2n + 1) reconstruc-
tions on Si(111) surface,” Phys. Rev. B 40, 1319 (1989).
Kolsky, H., Stress Waves in Solids (Oxford University Press, London, 1953).
Koplik, J., and H. Levine, “Interface moving through a random background,” Phys. Rev. B
32, 280 (1985).
Koschmieder, E.L., Bérnard Cells and Taylor Vortices (Cambridge University Press,
London, 1993).
Kosterlitz, J.M., and D.J. Thouless, “Ordering, metastability and phase transitions in two-
dimensional systems,” J. Phys. C 6, 1181 (1973).
Kramer, B., and A. MacKinnon, “Localization: theory and experiment,” Rep. Prog. Phys.
56, 1469 (1993).
Kuang, L., and H.J. Simon, “Diffusely scattered second harmonic generation from a silver
film due to surface plasmons,” Phys. Lett. A 197, 257 (1995).
Kulakhmetova, S.A., V.A. Saraikin, and L.I. Slepyan, “Plane problem of a crack in a lattice,”
Mech. Solids 19, 102 (1984).
Kulawansa, D.M., L.C. Jensen, S.C. Langford, and J.T. Dickinson, “Scanning electron
microscopy of the mirror region of silicate glass fracture surfaces,” J. Mater. Res. 9, 476
(1993).
Kun, F., A. Zapperi, and H.J. Herrmann, “Damage in fiber bundle models,” Eur. Phys. J. B
17, 269 (2000).
Kunin, I.A., Elastic Media with Microstructure (Springer, Berlin, 1982).
Kusy, R.P., and D.T. Turner, “Influence of the molecular weight of PMMA on fracture
surface energy in notched tension,” Polymer 17, 161 (1975).
Kusy, R.P., and D.T. Turner, “Influence of the molecular weight of poly(methyl
methacrylate) on fracture morphology in notched tension,” Polymer 18, 391 (1977).
Lagarkov, A.N., L.V. Panina, and A.K. Sarychev, “Effective magnetic permeability of com-
posite materials near the percolation threshold,” Sov. Phys. JETP 66, 123 (1987) [Zh.
Eskp. Teor. Fiz. 93, 215 (1987)].
Lamaignere, L., F. Carmona, and D. Sornette, “Experimental realization of critical thermal
fuse rupture,” Phys. Rev. Lett. 77, 2738 (1996).
Landau, L.D., and E.M. Lifshitz, Statistical Physics (Pergamon Press, London, 1980).
Langer, J.S., “Models of crack propagation,” Phys. Rev. A 46, 3123 (1992).
Langer, J.S., “Dynamic model of onset and propagation of fracture,” Phys. Rev. Lett. 70,
3592 (1993).
Langer, J.S., and A.E. Lobkovsky, “Critical examination of cohesive-zone models in the
theory of dynamic fracture,” J. Mech. Phys. Solids 46, 1521 (1998).
Langer, J.S., and H. Nakanishi, “Models of crack propagation. II. Two-dimensional model
with dissipation on the fracture surface,” Phys. Rev. E 48, 439 (1993).
Langer, J.S., and C. Tang, “Rupture propagation in a model of an earthquake fault,” Phys.
Rev. Lett. 67, 1043 (1991).
Langford, S.C., M. Zhenyi, and J.T. Dickinson, “Photon emission as a probe of chaotic
processes accompanying fracture,” J. Mater. Res. 4, 1272 (1989).
Larkin, A.I., and Yu.N. Ovchinnikov, “Pinning in type II superconductors,” J. Low Temp.
Phys. 34, 409 (1979).
Larralde, H., and R.C. Ball, “The shape of slowly growing cracks,” Europhys. Lett. 30, 87
(1995).
Larson, R.G., “Derivation of generalized Darcy equations for creeping flow in porous
media,” Ind. Eng. Chem. Fund. 20, 132 (1981).
610 References
Lawn, B., Fracture of Brittle Solids, 2nd ed. (Cambridge University Press, Cambridge,
1993).
Leath, P.L., and P.M. Duxbury, “Fracture of heterogeneous materials with continuous
distributions of local breaking strengths,” Phys. Rev. B 49, 14905 (1994).
Lee, H.-C., and M.E. Mear, “Effective properties of power-law solids containing elliptical
inhomogenities. I. Rigid inclusions,” Mech. Mater. 13, 313 (1992).
Lee, H.-C., and K.W. Yu, “Effective medium theory for strongly nonlinear composites:
comparison with numerical simulations,” Phys. Lett. A 197, 341 (1995).
Lemaire, E., Y. Ould Mohamed Abdelhaye, J. Larue, R. Benoit, P. Levitz, and H. Van
Damme, “Pattern formation in noncohesive and cohesive granular media,” Fractals 1,
968 (1993).
Leung, K.-t., and Z. Néda, “Pattern formation and selection in quasistatic fracture,” Phys.
Rev. Lett. 85, 662 (2000).
Levin, V.M., “Thermal expansion coefficients of heterogeneous materials,” Mekh. Tverd.
Tela 2, 83 (1967).
Levy, O., and D.J. Bergman, “The bulk effective response of non-linear random resistor
networks: numerical study and analytic approximations,” J. Phys.: Condens. Matter 5,
7095 (1993).
Levy, O., and D.J. Bergman, “Intrinsic optical bistability and resonances in nonlinear
composites,” Physica A 207 157 (1994a).
Levy, O., and D.J. Bergman, “Critical behavior of the weakly nonlinear conductivity and
flicker noise of two-component composites,” Phys. Rev. B 50, 3652 (1994b).
Levy, O., and D.J. Bergman, and D.G. Stroud, “Harmonic generation, induced nonlinearity,
and optical bistability in nonlinear composites,” Phys. Rev. E 52, 3184 (1995).
Levy, O., and R.V. Kohn, “Duality relations for non-Ohmic composites, with application
to behavior near percolation,” J. Stat. Phys. 90, 159 (1998).
Levy-Nathansohn, R., and D.J. Bergman, “Decoupling and testing of the generalized Ohm’s
law,” Phys. Rev. B 55 5425 (1997).
Li, G., P. Ponte Castañeda, and A.S. Douglas, “Constitutive models for ductile solids
reinforced by rigid spheroidal inclusions,” Mech. Mater 15, 279 (1993).
Li, R., and K. Sieradzki, “Ductile-brittle transition in random porous Au,” Phys. Rev. Lett.
68, 1168 (1992).
Li, W., R.K. Kalia, and P. Vashishta, “Amorphization and fracture in silicon diselenide
nanowires: a molecular dynamics study,” Phys. Rev. Lett. 77, 2241 (1996).
Li, X.-P., R.W. Nunes, and D. Vanderbilt, “Density-matrix electronic-structure method with
linear system-size scaling,” Phys. Rev. B 47, 10891 (1993).
Li, Y.S., and P.M. Duxbury, “Size and location of the largest current in a random resistor
network,” Phys. Rev. B 36, 5411 (1987).
Li, Y.S., and P.M. Duxbury, “Crack arrest by residual bonding in resistor and spring
networks,” Phys. Rev. B 38, 9257 (1988).
Liao, H.B., R.F. Xiao, H. Wang, K.S. Wong, and G.K.L. Wong, “Large third-order optical
nonlinearity in Au:TiO2 composite films measured on a femtosecond time scale,” Appl.
Phys. Lett. 72, 1817 (1998).
Lidorikis, E., M.E. Bachlechner, R.K. Kalia, A. Nakano, P. Vashishta, and G.Z. Voyiadjis,
“Coupling length scales for multiscale atomistic-continuum simulations: Atomistically
induced stress distribution in Si/Si3 N4 nanopixels,” Phys. Rev. Lett. 87, 086104-1 (2001)
Lim, S., Tsotsis, T.T., and Sahimi, M., “Molecular dynamics simulation of porous amorphous
carbon and transport of fluid mixtures therein,” J. Chem. Phys. (2003).
References 611
Lin, J.J., “Nonlinear I − V characteristics of the granular PrBa2 Cu3 O7−δ compound,” J.
Phys. Soc. Japan 61, 4125 (1992).
Lin, S.L., J. Mellor-Crumney, B.M. Pettitt, and G.N. Phillips, Jr., “Molecular dynamics on
a distributed-memory multiprocessor,” J. Comput. Phys. 13, 1022 (1992).
Lobb, C.J., P.M. Hui, and D. Stroud, “Nonuniversal breakdown behavior in superconducting
and dielectric composites,” Phys. Rev. B 36, 1956 (1987).
Lomdahl, P.S., D.M. Beazley, P. Tamayo, and N. Grφnbech-Jensen, “Multi-million particle
molecular dynamics on the CM-5,” Int. J. Mod. Phys. C 4, 1074 (1993).
López, J.M., and J. Schmittbuhl, “Anomalous scaling of fracture surfaces,” Phys. Rev. E
57, 6405 (1998).
Louis, E., and F. Guinea, “The fractal nature of fracture,” Europhys. Lett. 3, 871 (1987).
Lupkowski, M., and F. van Swol, “Ultrathin films under shear,” J. Chem. Phys. 95, 1995
(1995).
Ma, H., Xiao, R., and P. Sheng, “Third-order optical nonlinearity enhancement through
composite microstructures,” J. Opt. Soc. Am. B 15, 1022 (1998).
MacElroy, J.M.D., “Nonequilibrium molecular dynamics simulation of diffusion and flow
in thin microporous membranes,” J. Chem. Phys. 101, 5274 (1994).
Machová, A., “Dynamic microcrack initiation in alpha -iron,” Mater. Sci. Eng. A206, 279
(1996).
Machta, J., and R.A. Guyer, “Largest current in a random resistor network,” Phys. Rev. B
36, 2142 (1987).
Mahadevan, M., R.M. Bradley, and J.-M. Debierre, “Simulations of an electromigration-
induced edge instability in single-crystal metal lines,” Europhys. Lett. 45, 680 (1999).
Måløy, K.J., A. Hansen, E.L. Hinrichsen and S. Roux, “Experimental measurements of the
roughness of brittle cracks,” Phys. Rev. Lett. 68, 213 (1992).
Mandal, P.,A. Neumann,A.G. Jansen, P. Wyder, and R. Deltour, “Temperature and magnetic-
field dependence of the resistivity of carbon-black polymer composites,” Phys. Rev. B
55, 452 (1997).
Mandel, J., Plasticité Classique et Viscoplasticité CISM Udine Courses and Lectures, 97
(Springer, Berlin, 1972).
Mandelbrot, B.B., The Fractal Geometry of Nature (W.H. Freeman, San Francisco, 1982).
Mandelbrot, B.B., “Self-affine fractals and fractal dimension,” Physica Scripta 32, 257
(1985).
Mandelbrot, B.B., D.E. Passoja, and A.J. Paullay, “Fractal character of fracture surfaces of
metals,” Nature 308, 721 (1984).
Mandelbrot, B.B., and J.W. Van Ness, “Fractional Brownian motions, fractional noise and
applications,” SIAM Rev. 10, 422 (1968).
Manna, S.S., and B.K. Chakrabarti, “Dielectric breakdown in the presence of random
conductors,” Phys. Rev. B 36, 4078 (1987).
Manogg, P., “Investigation of the rupture of a Plexiglas plate by means of an optical method
involving high speed filming of the shadows originating around holes drilled in the plate,”
Int. J. Fract. 2, 604 (1966).
Mantese, J.W., W.I. Goldberg, D.H. Darling, H.G. Craighead, U.J. Gibson, R.A. Buhrman,
and W.W. Webb, “Excess low frequency conduction noise in a granular composite,” Solid
State Commun. 37, 353 (1981).
Marcellini, P., “Periodic Solutions and homogenization of nonlinear variational problems,”
Ann. Mat. Pura. Appl. 4, 139 (1978).
Marder, M., “New dynamical equation for cracks,” Phys. Rev. Lett. 66, 2484 (1991).
Marder, M., “Statistical mechanics of cracks,” Phys. Rev. E 54, 3442 (1996).
612 References
Marder, M., and J. Fineberg, “How things break,” Phys. Today 49 (No. 9), 24 (1996).
Marder, M., and S.P. Gross, “Origin of crack tip instabilities,” J. Mech. Phys. Solids 43, 1
(1995).
Marder, M., and X. Liu, “Instability in lattice fracture,” Phys. Rev. Lett. 71, 2417 (1993).
Markel, V.A., V.M. Shalaev, E.B. Stechel, W. Kim, and R.L. Armstrong, “Small-particle
composites. I. Linear optical properties,” Phys. Rev. B 53, 2425 (1996).
Markel, V.A., V.M. Shalaev, P. Zhang, W. Huynh, L. Tay, T.L. Haslett, and M. Moskovits,
“Near-field optical spectroscopy of individual surface-plasmon modes in colloid
clusters,” Phys. Rev. B 59, 10903 (1999).
Marks, N.A., D.R. McKenzie, B.A. Paithorpe, M. Bernasconi, and M. Parrinello, “Ab initio
simulations of tetrahedral amorphous carbon,” Phys. Rev. B 54, 9703 (1996).
Martin, J.E., and M.B. Heaney, “Reversible thermal fusing model of carbon black current-
limiting thermistors,” Phys. Rev. B 62, 9390 (2000).
Mart́in, T., P. Español, M.A. Rubio, and I. Zúñiga, “Dynamic fracture in a discrete model
of a brittle elastic solid,” Phys. Rev. E 61, 6120 (2000).
Martins, J.L., and A. Zunger, “Bond lengths around isovalent impurities and in
semiconductor solid solutions,” Phys. Rev. B 30, 6217 (1984).
Martyna, G.J., M.L. Klein, and M. Tuckerman, “Nose-Hoover chains: the canonical
ensemble via continuous dynamics,” J. Chem. Phys. 97, 2635 (1992).
Martyna, G.J., D.T. Tobias, and M.L. Klein, “Constant pressure molecular dynamics
algorithms,” J. Chem. Phys. 101, 4177 (1994).
Matthews, J.W., andA.E. Blakeslee, “Defects in epitaxial multilayers. I. Misfit dislocations,”
J. Cryst. Growth 32, 265 (1974).
Mauri, F., G. Galli, and R. Car, “Orbital formulation for electronic-structure calculations
with linear system-size scaling,” Phys. Rev. B, 47, 9973 (1993).
McAnulty, P., L.V. Meisel, and P.J. Cote, “Hyperbolic distributions and fractal character of
fracture surfaces,” Phys. Rev. A 46, 3523 (1992).
Meakin, P., “A simple model for elastic fracture in thin films,” Thin Solid Films 151, 165
(1987).
Meakin, P., “Simple kinetic models for material failure and deformation,” Herrmann and
Roux (1990), p. 291.
Meakin, P., Fractals, Scaling and Growth far from Equilibrium (Combridge University
Press, Cambridge, 1998).
Meakin, P., G. Li, L.M. Sander, E. Louis, and F. Guinea, “A simple two-dimensional model
for crack propagation", J. Phys. A 22, 1393 D (1989).
Mecholsky, J.J., in Strength of Inorganic Glass, edited by C.R. Kurkjian (Plenum, New
York, 1985).
Mecholsky, J.J., T.J. Mackin, and D.E. Passoja, Adv. Ceramics 22, 127 (1988).
Mecholsky, J.J., D.E. Passoja, and K.S. Feinberg-Ringel, “Quantitative analysis of brittle
fracture surfaces using fractal geometry,” J. Am. Ceram. Soc. 72, 60 (1989).
Meek, J.M., and J.D. Craggs, Electrical Breakdown of Gases (Wiley, New York, 1978).
Mehrabi, A.R., H., Rassamdana, and M., Sahimi, “Characterization of long-range
correlations in complex distributions and profiles,” Phys. Rev. E 56, 712 (1997).
Mehrabi, A.R., and M. Sahimi, “Diffusion of ionic particles in charged disordered media,”
Phys. Rev. Lett. 82, 735 (1999).
Meir, Y., R. Blumenfeld, A. Aharony, and A.B. Harris, “Series analysis of randomly diluted
nonlinear resistor networks,” Phys. Rev. B 34, 3424 (1986).
Mel’cuk, A.I., R.C. Giles, and H. Gould, “Molecular dynamics simulation of liquids on the
Connection Machine,” Computers in Physics, 311 (May/June 1991).
References 613
Mercer, J.L., “Tight-binding models for compounds: Application to SiC,” Phys. Rev. B 54,
4650 (1996).
Mermin, N.D., “Thermal properties of the inhomogeneous electron gas,” Phys. Rev. 137,
A1441 (1965).
Michel, J.C., “A self-consistent estimate of the potential of a composite made of two power-
law phases with the same exponent,” C. R. Acad. Sci. Paris Ser. II 322, 447 (1996).
Mikitishin, S.I., Y.-N. Skhonitskii, and A.N. Tynnyi, Fiz. Khim. Mekh. Mater. 5, 69 (1969).
Miksis, M.J., “Effective dielectric constant of a nonlinear composite material,” SIAM J.
Appl. Math. 43, 1140 (1983).
Miller, S., and R. Reifenberger, “Improved method for fractal analysis using scanning probe
microscopy,” J. Vac. Sci. Technol. B10, 1203 (1992).
Milman, V.Y., “Fracture surfaces: A critical review of fractal studied and a novel morpho-
logical analysis of scanning tunneling microscopy measurements,” Prog. Mater. Sci. 38,
425 (1994).
Milman, V.Y., R. Blumenfeld, N.A. Stelmashenko, and R.C. Ball, “Comment on ’Exper-
imental measurements of the roughness of brittle cracks’,” Phys. Rev. Lett. 71, 204
(1993).
Milman, V.Y., N.A. Stelmashenko, and R. Blumenfeld, “Fracture surfaces: A critical review
of fractal studies and a novel morphological analysis of scanning tunneling microscopy
measurements,” Prog. Mater. Sci. 38, 425 (1994).
Milstein, F., in Mechanics of Solids, edited by H.G. Hopkins and M.J. Sewell (Pergamon
Press, Oxford, 1982).
Milton, G.W., “Bounds on the electromagnetic, elastic, and other properties of two-
component composites,” Phys. Rev. Lett. 46, 542 (1981a).
Milton, G.W., “Bounds on the elastic and transport properties of two-component
composites,” J. Mech. Phys. Solids 30, 177 (1981b).
Milton, G.W., “Bounds on the elastic and transport properties of two-component
composites,” J. Mech. Phys. Solids 30, 177 (1982).
Milton, G., “Modeling the properties of composites by laminates,” in Homogenization and
Effective Moduli of Materials and Media, edited by J.L. Ericksen, D. Linderlehrer, R.
Kohn, and J.-L. Lions (Springer, Berlin, 1986), p. 150.
Mistriotis, A.D., N. Flytzanis, and S.C. Farantos, “Potential model for silicon clusters,”
Phys. Rev. B 39, 1212 (1989).
Mitchell, M.W., and D.A. Bonnell, “Quantitative topographic analysis of fractal surfaces
by scanning tunneling microscopy,” J. Mater. Res. 5, 2244 (1990).
Molinari, A., G.R. Canova, and S. Ahzi, “A self-consistent approach of the large deformation
polycrystal viscoplasticity,” Acta Metall. Mater. 35, 2983 (1987).
Morales, J.J., and M.J. Nuevo, “Comparison of link-cell and neighborhood tables on a range
of computers,” Comput. Phys. Commun. 69, 223 (1992).
Morel, S., J. Schmittbuhl, E. Bouchaud, and G. Valentin, “Scaling of crack surfaces and
implications for fracture mechanics,” Phys. Rev. Lett. 85, 1678 (2000).
Morel, S., J. Schmittbuhl, J.M. López, and G. Valentin, Phys. Rev. E 58, 6999 (1998).
Moreno, Y., J.B. Gómez, and A.F. Pacheco, “Fracture and second-order phase transitions,”
Phys. Rev. Lett. 85, 2865 (2000).
Mori, Y., K. Kaneko, and M. Wadati, “Fracture dynamics by quenching. I. Crack patterns,”
J. Phys. Soc. Japan 60, 1591 (1991).
Morrissey, J.W., and J.R. Rice, “Crack front waves,” J. Mech. Phys. Solids 46, 467 (1998).
Moskovitz, M., “Surface-enhanced spectroscopy,” Rev. Mod. Phys. 57, 783 (1985).
614 References
Mosolov, A.B., “Mechanics of fractal cracks in brittle solids,” Europhys. Lett. 24, 673
(1993).
Mott, N.F., “Brittle fracture in mild steel plates,” Engineering 165, 16 (1948).
Moulinec, H., and P. Suquet, “A FTT-based numerical method for computing the mechanical
properties of composites from images of their microstructure,” in Microstructure-
Properly Interactions in Composite Materials, edited by R. Pyrz (Kluwer, The
Netherlands, 1995), p. 235.
Movchan, A.B., and J.R. Willis, “Dynamic weight functions for a moving crack. II. Shear
loading,” J. Mech. Phys. Solids 43, 1369 (1995).
Mu, Z.Q., and C.W. Lung, “Studies on the fractal dimension and fracture toughness of
steel,” J. Phys. D 21, 848 (1988).
Müller, K., B. Mehlig, F. Milde, and M. Schreiber, “Statistics of wave functions in disordered
and in classically chaotic systems,” Phys. Rev. Lett. 78, 215 (1997).
Mura, D., L. Colombo, R. Bertoncini, and G. Mula, “Structure and chemical order of bulk
Si1−x Cx amorphous alloys,” Phys. Rev. B 58, 10357 (1998).
Murray, R.T., C.J. Keily, and M. Hopkinson, “Crack formation in III-V epilayers grown
under tensile strain on InP(001) substrates,” Philos. Mag. 74, 383 (1996).
Murthy, C.S., S.F. O’Shea, and I.R. McDonald, “Electrostatic interactions in molecular
crystals. Lattice dynamics of solid nitrogen and carbon dioxide,” Mol. Phys. 50, 531
(1983).
Muskhelishvili, N.I., Some Basic Problems of the Mathematical Theory of Elasticity (P.
Noordhoff, Groningen, Holland, 1953).
Nakamura, T., and D.M. Parks, “Three-dimensional stress field near the crack front of a
thin elastic plate,” J. Appl. Mech. 55, 805 (1988).
Nakano, A., R.K. Kalia, and P. Vashishta, “Growth of pore interfaces and roughness of
fracture surfaces in porous silica: million particle molecular-dynamics simulations,”
Phys. Rev. Lett. 73, 2336 (1994).
Nakano, A., R.K. Kalia, and P. Vashishta, “Dynamics and morphology of brittle cracks: a
molecular-dynamics study of silicon nitride,” Phys. Rev. Lett. 75, 3138 (1995).
Namgoong, E., and J.S. Chun, “The effect of ultrasonic vibration on hard chromium plating
in a modified selfregulating high speed bath,” Thin Solid Films 120, 153 (1984).
Narayan, O., and D.S. Fisher, “Dynamics of sliding charge-density waves in 4 −
dimensions,” Phys. Rev. Lett. 68, 3615 (1992).
Narayan, O., and D.S. Fisher, “Nonlinear fluid flow in random media: critical phenomena
near threshold,” Phys. Rev. B 49, 9469 (1994).
Nelson, D.R., and B.I. Halperin, “Dislocation-mediated melting in two dimensions,” Phys.
Rev. B 19, 2457 (1979).
Nie, S., and S.R. Emory, “Probing single molecules and single nanoparticles by surface-
enhanced Raman scattering,” Science 275, 1102 (1997).
Niemeyer, L., L. Pietronero, and H.J. Wiesmann, “Fractal dimension of dielectric
breakdown,” Phys. Rev. Lett. 52, 1033 (1984).
Nishioka, K., and J.K. Lee, “Temperature dependence of the ideal fracture strength of a
b.c.c. crystal,” Philos. Mag. A 44, 779 (1981).
Nishioka, K., S. Nakamura, T. Shimamoto, and H. Fujiwara, “Lattice instability theory of
fracture,” Scripta Mettal. 14, 497 (1980).
Noble, B., Methods Based on the Wiener–Hopf Technique for the Solution of Partial
Differential Equations (Pergamon, New York, 1958).
Noble, B., and J.W. Daniel, Applied Linear Algebra, 2nd ed. (Prentice-Hall, Englewood
Cliffs, NJ, 1977).
References 615
Nosé, S., “A unified formulation of the constant temperature molecular dynamics methods,”
J. Chem. Phys. 81, 511 (1984).
Noskov, M.D., V.R. Kukhta, and V.V. Lopatin, “Simulation of the electrical discharge
development in inhomogeneous insulators,” J. Phys. D 28, 1187 (1995).
Nyikos, L., and T. Pajkossy, “Short communication—Diffusion to fractal surfaces,”
Electrochim. Acta 31, 1347 (1985).
Obukhov, S.P., “The problem of directed percolation,” Physica A 101, 145 (1980).
O’Dwyer, J.J., The Theory of Electrical Conduction and Breakdown in Solid Dielectrics
(Clarendon Press, Oxford, 1973).
Ohring, M., Reliability and Failure of Electronic Materials and Devices (Academic Press,
San Diego, 1998).
Olsen, K.B., R. Madariaga, and R.J. Archuleta, “Three-dimensional dynamic simulation of
the 1992 Landers earthquake,” Science 278, 834 (1997).
Olson, T., “Improvements on Taylor’s upper bound for rigid-plastic composites,” Mater.
Sci. Eng. A 175, 15 (1994).
Omeltchenko, A., A. Nakano, R.K. Kalia, and P. Vashishta, “Structure, mechanical
properties, and thermal transport in microporous silicon nitride-molecular-dynamics
simulations on a parallel machine,” Europhys. Lett. 33, 667 (1996).
Omeltchenko, A., J. Yu, R.K. Kalia, and P. Vashishta, “Crack front propagation and fracture
in a graphite sheet: a molecular dynamics study on parallel computers,” Phys. Rev. Lett.
78, 2148 (1997).
Ordejón, P., D.A. Drabold, R.M. Martin, and M.P. Grumbach, “Linear system-size scaling
methods for electronic-structure calculations,” Phys. Rev. B 51, 1456 (1995).
Orowan, E., “Crystal plasticity—III. On the mechanism of the glide process,” Z. Phys. 89,
605 (1934).
Orowan, E., “Energy criteria of fracture,” Weld. Res. Supp. 34, 157 (1955).
Ortiz, M., and R. Phillips, “Nanomechanics of defects in solids,” Adv. Appl. Mech. 36, 1
(1999).
Pande, C.S., L.E. Richards, N. Louat, B.D. Dempsey, and A.J. Schwoeble, “Fractal
characterization of fractured surfaces,” Acta Metall. 35, 1633 (1987).
Pannetta, C., L. Reggiani, and Gy. Trefán, “Scaling and universality in electrical failure of
thin films,” Phys. Rev. Lett. 84, 5006 (2000).
Parleton, L.G., “Determination of the growth of branched cracks by numerical methods,”
Eng. Fract. Mech. 11, 343 (1979).
Parr, R.G., and W. Yang, Density Functional Theory of Atoms and Molecules (Oxford
University Press, New York, 1989).
Parrinello, M., and A. Rahman, “Polymorphic transitions in single crystals: A new molecular
dynamics method,” J. Appl. Phys. 52, 7182 (1981).
Paskin, P., A. Gohar, and G.J. Dienes, “Computer simulation of crack propagation,” Phys.
Rev. Lett. 44, 940 (1980).
Paskin, P., D.K. Som, and G.J. Dienes, “Computer simulation of crack propagation: lattice
trapping,” J. Phys. C 14, L171 (1981).
Passoja, D.E., Adv. Ceramics 22, 101 (1988).
Passoja, D.E., and D.J. Amborski, Microstruct. Sci. 6, 143 (1978).
Payne, M.C., M.P. Teter, D. Allen, T.A. Arias, and J.D. Joannopoulos, “Iterative minimiza-
tion techniques for ab initio total-energy calculations: molecular dynamics and conjugate
gradients,” Rev. Mod. Phys. 64, 1045 (1992).
Pederson, M.R., and K.A. Jackson, “Pseudoenergies for simulations on metallic systems,”
Phys. Rev. B 43, 7312 (1991).
616 References
Ponte Castañeda, P., and G. deBotton, “On the homogenized yield strength of two-phase
composites,” Proc. R. Soc. Lond. A 438, 419 (1992).
Ponte Castañeda, P., G. deBotton, and G. Li, “Effective properties of nonlinear inhomoge-
neous dielectrics,” Phys. Rev. B 46, 4387 (1992).
Ponte Castañeda, P., and M. Kailasam, “Nonlinear electrical conductivity in heterogeneous
media,” Proc. R. Soc. Lond. A 453, 793 (1997); 453, 1791(E).
Ponte Castañeda, P., and M.V. Nebozhyn, “Exact second-order estimates of the self-
consistent type for nonlinear composite materials,” Mech. Mater. (1997).
Ponte Castañeda, P., and P. Suquet, “On the effective mechanical behavior of weakly
inhomogeneous nonlinear materials,” Europ. J. Mech. A Solids 14, 205 (1995).
Ponte Castañeda, P., and P. Suquet, “Nonlinear composites,” Adv. Appl. Mech. 34 (Academic
Press, San Diego, 1998), p. 172.
Ponte Castañeda, P., and J.R. Willis, “On the overall properties of nonlinearly viscous
composites,” Proc. R. Soc. Lond. A 416, 217 (1988).
Ponte Castañeda, P., and J.R. Willis, “The effective behavior of nonlinear composites:
A comparison between two methods,” in Continuum Models and Discrete Systems
(CMDS 7), edited by K.H. Anthony and H.-H. Wagner (Trans Tech, Aedermannsdorf,
Switzerland, 1993), p. 351.
Ponte Castañeda, P., and J.R. Willis, “The effect of spatial distribution on the effective
behavior of composite materials and cracked media,” J. Mech. Phys. Solids 43, 1919
(1995).
Poon, C.Y., R.S. Sayles, and T.A. Jones, “Surface measurement and fractal characterization
of naturally fractured rocks,” J. Appl. Phys. 25, 1269 (1992).
Porto, M., S. Havlin, S. Schwarzer, and A. Bunde, “Optimal path in strong disorder and
shortest path in invasion percolation with trapping,” Phys. Rev. Lett. 79, 4060 (1997).
Prakash, V., and R.J. Clifton, “Experimental and analytical investigation of dynamic fracture
under conditions of plane strain,” in Fracture Mechanics: 22nd Symposium, vol. 1,
edited by H.A. Ernst, A. Saxena, and D.L. McDowell (American Society for Testing and
Materials, Philadelphia, 1992), p. 412.
Pratt, A.K., and P.L. Green, “Measurement of the dynamic fracture toughness of
polymethylmethacrylate by high-speed photography,” Eng. Fract. Mech. 6, 71 (1974).
Pulay, P., “Ab initio calculation of force constants and equilibrium geometries in polyatomic
molecules. I. Theory,” Mol. Phys. 17, 197 (1969).
Qiu, S.-Y., C.Z. Wang, K.M. Ho, and C.T. Chan, “Tight-binding molecular dynamics with
linear system-size scaling,” J. Phys.: Condens. Matter 6, 9153 (1994).
Qiu, Y.P., and G.J. Weng, “A theory of plasticity for porous materials and particle-reinforced
composites,” J. Appl. Mech. 59, 261 (1992).
Rahman, A., “Correlation in the motion of atoms in liquid argon,” Phys. Rev. 136, A405
(1964).
Rahman, A., and F.H. Stillinger, “Molecular dynamics study of liquid water,” J. Chem.
Phys. 55, 3336 (1971).
Raine, A.R.C., D. Fincham, and W. Smith, “Systolic loop methods for molecular dynamics
simulation using multiple transputers,” Comput. Phys. Commun. 55, 13 (1989).
Räisänen, V.I., M.J. Alava, K.J. Niskanen, and R.M. Nieminen, “Does the shear-lag model
apply to random fiber networks?” J. Mater. Res. 12, 2725 (1997).
Räisänen, V.I., E.T. Seppala, M.J.Alava, and P.M. Duxbury, “Quasistatic cracks and minimal
energy surfaces,” Phys. Rev. Lett. 80, 329 (1998).
Rambaldi, S., and O. Pinazza, “An accurate fractional Brownian motion generator,” Physica
A 208, 21 (1994).
618 References
Ramanathan, S., and D.S. Fisher, “Dynamics and instabilities of planar tensile cracks in
heterogeneous media,” Phys. Rev. Lett. 79, 877 (1997).
Ramanathan, S., and D.S. Fisher, “Onset of propagation of planar cracks in heterogeneous
media,” Phys. Rev. B 58, 6026 (1998).
Rammal, R., and A.-M.S. Tremblay, “Resistance noise in nonlinear resistor networks,”
Phys. Rev. Lett. 58, 415 (1987).
Ramulu, M., and A.S. Kobayashi, “Mechanics of crack curving and branching—a dynamic
fracture analysis,” Int. J. Fract. 27, 187 (1985).
Ramulu, M., and A.S. Kobayashi, “Strain energy density criteria for dynamic fracture and
dynamic crack branching,” Theore. Appl. Fract. Mech. 5, 117 (1986).
Rapaport, D.C., “Multi-million particle molecular dynamics. III. Design consideration for
data-parallel processing,” Comput. Phys. Commun. 76, 301 (1993).
Rapaport, D.C., The Art of Molecular Dynamics (Cambridge University Press, London,
1995).
Rappe, A., and J.D. Joannopoulos, in Computer Simulation in Materials Science, edited by
M. Meyer and V. Pontikis, NATO ASI Vol. 205, p. 409 (1991).
Rautiainen, T.T., M.J. Alava, and K. Kaski, “Dynamics of fracture in dissipative systems,”
Phys. Rev. E 51, R2727 (1998).
Ravi-Chandar, K., and W.G. Knauss, “Dynamic crack-tip stress under stress wave loading—
a comparison of theory and experiment,” Int. J. Fract. 20, 209 (1982).
Ravi-Chandar, K., and W.G. Knauss, “An experimental investigation into dynamic fracture.
I. Crack initiation and arrest,” Int. J. Fract. 25, 247 (1984a).
Ravi-Chandar, K., and W.G. Knauss, “An experimental investigation into dynamic fracture:
II. Microstructural aspects,” Int. J. Fract. 26, 65 (1984b).
Ravi-Chandar, K., and W.G. Knauss, “An experimental investigation into dynamic frac-
ture: III. On steady-state crack propagation and crack branching,” Int. J. Fract. 26, 141
(1984c).
Ravi-Chandar, K., and B. Yang, “On the role of microcracks in the dynamic fracture of
brittle materials,” J. Mech. Phys. Solids 45, 535 (1997).
Ray J.R., “Elastic constants and statistical ensembles in molecular dynamics,” Comput.
Phys. Rep. 8, 109 (1988).
Ray, P., and B.K. Chakrabarti, “The critical behaviour of fracture properties of dilute brittle
solids near the percolation threshold,” J. Phys. C 18, L185 (1985a).
Ray, P., and B.K. Chakrabarti, “A microscopic approach to the statistical fracture analysis
of disordered brittle solids,” Solid State Commun. 53, 477 (1985b).
Ray, P., and B.K. Chakrabarti, “Strength of disordered solids,” Phys. Rev. B 38, 715 (1988).
Reidle, J., P. Gumbsch, H.F. Fischmeister, V.G. Glebovsky, and V.N. Semenov, “Fracture
studies of tungsten single crystals,” Mater. Lett. 20, 311 (1994).
Reuss, A., “Calculation of the flow limits of mixed crystals on the basis of the plasticity of
the monocrystals,” Z. Angew. Math. Mech. 9, 49 (1929).
Rice, J.R., “Mathematical analysis in the mechanics of fracture,” in Fracture: An Advanced
Treatise, Vol. II, edited by H. Liebowitz (Academic Press, New York, 1968), p. 191.
Rice, J.R., “On the structure of stress-strain relations for time-dependent plastic deformation
in metals,” J. Appl. Mech. 37, 728 (1970).
Rice, J.R., and G.E. Beltz, “The activation energy for dislocation nucleation at a crack,” J.
Mech. Phys. Solids 42, 333 (1994).
Rice, J.R., Y. Ben-Zion, and K.S. Kim, “Three dimensional perturbation solution for a
dynamic planar crack moving unsteadily in a model elastic solid,” J. Mech. Phys. Solids
42, 813 (1994).
References 619
Rice, J.R., and R. Thomson, “Ductile versus brittle behaviour of crystals,” Philos. Mag. 29,
73 (1974).
Robertson, D.H., D.W. Brenner, and J.W. Mintmire, “Energetics of nanoscale graphitic
tubules,” Phys. Rev. B 45, 12592 (1992).
Robertson, M.C., C.G. Sammis, M. Sahimi, and A.J. Martin, “Fractal analysis of three-
dimensional spatial distributions of earthquakes with a percolation interpretation,” J.
Geophys. Res. 100B, 609 (1995).
Rodbell, K.P., M.V. Rodriguez, and P.J. Ficalora, “The kinetics of electromigration,” J.
Appl. Phys. 61, 2844 (1987).
Rodgers, S.T., and K.F. Jensen, “Multiscale modeling of chemical vapor deposition,” J.
Appl. Phys. 83, 524 (1998).
Ronsin, O., F. Heslot, and B. Perrin, “Experimental study of quasistatic brittle crack
propagation,” Phys. Rev. Lett. 75, 2252 (1995).
Rosakis, A.J., J. Duffy, and L.B. Freund, “The determination of dynamic fracture toughness
of AISI 4340 steel by the shadow spot method,” J. Mech. Phys. Solids 32, 443 (1984).
Rosakis, A.J., and L.B. Freund, “The effect of crack tip plasticity on the determination of
dynamic stress intensity factors by the optical method of caustics,” J. Appl. Mech. 48,
302 (1981).
Rosakis, A.J., O. Samudrala, and D. Coker, “Cracks faster than the shear wave speed,”
Science 284, 1337 (1999).
Rosakis, A.J., and A.T. Zehnder, “On the dynamic fracture of structural metals,” Int. J.
Fract. 27, 169 (1985).
Roth, J., F. Gähler, and H.-R. Trebin, “A molecular dynamics run with 5180116000
particles,” Int. J. Mod. Phys. C 11, 317 (2000).
Roux, S., and D. Francois, “A simple model for ductile fracture of porous materials,” Scripta
Metall. 25, 1087 (1991).
Roux, S., A. Hansen, and E. Guyon, “Criticality in non-linear transport properties of
heterogeneous materials,” J. Physique 48, 2125 (1987).
Roux, S., A. Hansen, H.J. Herrmann, and E. Guyon, “Rupture of heterogeneous media in
the limit of infinite disorder,” J. Stat. Phys. 52, 251 (1988).
Roux, S., and H.J. Herrmann, “Disorder-induced nonlinear conductivity,” Europhys. Lett.
4, 1227 (1987).
Rudd, E., and J.Q. Broughton, “Coarse-grained molecular dynamics and the atomic limit
of finite elements,” Phys. Rev. B 58, 5893 (1998).
Rundle, J.B., and W. Klein, “Nonclassical nucleation and growth of cohesive tensile cracks,”
Phys. Rev. Lett. 63, 171 (1989).
Runge, E., and E.K.U. Gross, “Density-functional theory for time-dependent systems,”
Phys. Rev. Lett. 52, 997 (1984).
Ryckaert, J.P., G. Ciccotti, and H.J.C. Berendsen, “Numerical integration of the cartesian
equations of motion of a system with constraints: Molecular dynamics of n-alkane,” J.
Comput. Phys. 23, 327 (1977).
Ryckaert, J.P., “Special geometrical constraints in the molecular dynamics of chain
molecules,” Mol. Phys. 55, 549 (1985).
Sachs, G., “Zur Ableitung einer Fleissbedingun,” Z. Ver. Dtsch. Ing. 72, 734 (1928).
Safonov, V.P., V.M. Shalaev, V.A. Markel, Y.E. Danilova, N.N. Lepeshkin, W. Kim, S.G.
Rautian, and R.L. Armstrong, “Spectral dependence of selective photomodification in
fractal aggregates of colloidal particles,” Phys. Rev. Lett. 80, 1102 (1998).
Sahimi, M., “Nonlinear transport processes in disordered media,” AIChE J. 39, 369 (1993a).
620 References
Sahimi, M., “Flow phenomena in rocks: From continuum models to fractals, percolation,
cellular automata and simulated annealing,” Rev. Mod. Phys. 65, 1395 (1993b).
Sahimi, M., Applications of Percolation Theory (Taylor and Francis, London, 1994a).
Sahimi, M., “Long-range correlated percolation and flow and transport in heterogeneous
porous media,” J. Physique I France 4, 1263 (1994b).
Sahimi, M., “Effect of long-range correlations on transport phenomena in disordered
media,” AIChE J. 41, 229 (1995a).
Sahimi, M., Flow and Transport in Porous Media and Fractured Rock (VCH, Weinheim,
Germany, 1995b).
Sahimi, M., “Non-linear and non-local transport processes in heterogeneous media: From
long-range correlated percolation to fracture and materials breakdown,” Phys. Rep. 306,
295 (1998).
Sahimi, M., and S. Arbabi, “Force distribution, multiscaling, and fluctuations in disordered
elastic media,” Phys. Rev. B 40, 4975 (1989).
Sahimi, M., and S. Arbabi, “On correction to scaling for two- and three-dimensional scalar
and vector percolation,” J. Stat. Phys. 62, 453 (1991).
Sahimi, M., and S. Arbabi, “Percolation and fracture in disordered solids and granular
media: Approach to a fixed point,” Phys. Rev. Lett. 68, 608 (1992).
Sahimi, M., and S. Arbabi, “Mechanics of disordered solids. III. Fracture properties,” Phys.
Rev. B 47, 713 (1993).
Sahimi, M., and S. Arbabi, “Scaling laws for fracture of heterogeneous materials and rock,”
Phys. Rev. Lett. 77, 3689 (1996).
Sahimi, M., and J.D. Goddard, “Elastic percolation models for cohesive mechanical failure
in heterogeneous systems,” Phys. Rev. B 33, 7848 (1986).
Sahimi, M., M.C. Robertson, and C.G. Sammis, “Fractal distribution of earthquake hypocen-
ters and its relation with fault patterns and percolation,” Phys. Rev. Lett. 70, 2186
(1993).
Sammonds, P.R., P.G. Meredith, M.R. Ayling, and S.A. Murell, in Fracture of Concrete and
Rock, edited by S.P. Shah, S.E. Swartz, and B. Barr (Elsevier, London, 1989), p. 101.
Sánchez, A., F. Guinea, E. Louis, and V. Hakim, “On the fractal characteristics of the eta
model,” Physica A 191, 123 (1992).
Sanchez-Palancia, E., Non-homogeneous Media and Vibration Theory (Lecture notes in
Physics) 127 (Springer, Heidelberg, 1980), p. 12.
Sanz-Serna, J.M., “Symplectic integration for Hamiltonian systems: An overview,” Acta
Numerica 1, 243 (1992).
Sarychev, A.K., D.J. Bergman, and Y. Yagil, “Optical and microwave properties of metal-
insulator thin films: possibility of light localization,” Physica A 207, 372 (1994).
Sarychev,A.K., D.J. Bergman, and Y. Yagil, “Theory of the optical and microwave properties
of metal-dielectric films,” Phys. Rev. B 51, 5366 (1995).
Sarychev, A.K., and V.M. Shalaev, “Electromagnetic field fluctuations and optical
nonlinearities in metal-dielectric composites,” Phys. Rep. 335, 275 (2000).
Sarychev, A.K., and V.M. Shalaev, “Giant high-order field moments in metal-dielectric
composites,” Physica A 266, 115 (1999).
Sarychev, A.K., V.A. Shubin, and V.M. Shalaev, “Anderson localization of surface plasmons
and nonlinear optics of metal-dielectric composites,” Phys. Rev. B 60, 16389 (1999).
Sawada, Y., S. Ohta, M. Yamazaki, and H. Honjo, “Self-similarity and a phase-transition-
like behavior of a random growing structure governed by a nonequilibrium parameter,”
Phys. Rev. A 26, 3557 (1982).
References 621
Schardin, H., “Velocity effects in fracture,” in Fracture, edited by B.L. Averbach (MIT
Press, Cambridge, MA, 1959), p. 297.
Schimschak, M., and J. Krug, “Electromigration-induced breakup of two-dimensional
voids,” Phys. Rev. Lett. 80, 1674 (1998).
Schmittbuhl, J., S., Gentier, and S. Roux, S., “Field measurements of the roughness of fault
surfaces,” Geophys. Res. Lett. 20, 639 (1993).
Schmittbuhl, J., S. Roux, J.-P. Vilotte, and K.J. Måløy, “Interfacial crack pinning: effect of
nonlocal interactions,” Phys. Rev. Lett. 74, 1787 (1995).
Schöen, M., “Structure of a simple molecular dynamics. II. Design considerations for
distributed processing,” Comput. Phys. Commun. 52, 175 (1989).
Schwartz, L.M., S. Feng, M.F. Thorpe, and P.N. Sen, “Behavior of depleted elastic networks:
Comparison of effective-medium and numerical calculations,” Phys. Rev. B 32, 4607
(1985).
Scott, I.G., “Basic Acoustic Emission,” in Nondestructive Testing Monographs and Tracts,
Vol. 6 (Gordon and Breach New York, 1991), p. 124.
Shalaev, V.M., R. Botet, and A.V. Butenko, “Localization of collective dipole excitations
on fractals,” Phys. Rev. B 48, 6662 (1993).
Shalaev, V.M., R. Botet, J. Mercer, and E.B. Stechel, “Optical properties of self-affine thin
films,” Phys. Rev. B 54, 8235 (1996a).
Shalaev, V.M., C. Douketis, T. Haslett, T. Stuckless, and M. Moskovits, “Two-photon elec-
tron emission from smooth and rough metal films in the threshold region,” Phys. Rev. B
53, 11193 (1996b).
Shalaev, V.M., V.A. Markel, E.Y. Poliakov, R.L. Armstrong, V.P. Safonov, A.K. Sarychev,
“Nonlinear optical phenomena in nanostructured fractal materials,” J. Nonlin. Optic.
Phys. & Mat. 7, 131 (1998).
Shalaev, V.M., and A.K. Sarychev, “Nonlinear optics of random metal-dielectric films,”
Phys. Rev. B 57, 13265 (1998).
Sharon, E., and J. Fineberg, “Microbranching instability and the dynamic fracture of brittle
materials,” Phys. Rev. B 54, 7128 (1996).
Sharon, E., and J. Fineberg, “Universal features of the micro-branching instability in
dynamic fracture,” Philos. Mag. B 78, 243 (1998).
Sharon, E., and J. Fineberg, “On massless cracks and the continuum theory of fracture,”
Nature 397, 333 (1999).
Sharon, E., S.P. Gross, and J. Fineberg, “Local crack branching as a mechanism for instability
in dynamic fracture,” Phys. Rev. Lett. 74, 5096 (1995).
Sharon, E., S.P. Gross, and J. Fineberg, “Energy dissipation in dynamic fracture,” Phys.
Rev. Lett. 76, 2117 (1996).
Sharp, S.J., M.F. Ashby, and N.A. Fleck, “Material response under static and sliding
indentation loads,” Acta Met. 41, 685 (1993).
Shenoy, V.B., R. Miller, E.B. Tadmor, R. Phillips, and M. Ortiz, “Quasicontinuum models
of interfacial structure and deformation,” Phys. Rev. Lett. 80, 742 (1998).
Shenoy, V.B., R. Miller, E.B. Tadmor, D. Rodney, and R. Phillips, “An adaptive finite
element approach to atomic-scale mechanics—the quasicontinuum method,” J. Mech.
Phys. Solids 47, 611 (1999).
Shenoy, V.B., R. Phillips, and E.B. Tadmor, “Nucleation of dislocation beneath a plane
strain indenter,” J. Mech. Phys. Solids 48, 649 (2000).
Shimojo, F., I. Ebbsjö, R.K. Kalia, A. Nakano, J.P. Rino, and P. Vashishta, “Molecular
dynamics simulation of structural transformation in silicon carbide under pressure,”
Phys. Rev. Lett. 84, 3338 (2000).
622 References
Shinkai, N., “Fracture and fractography of flat glass,” in Fractography of Glass, edited by
R.C. Bradt and R.E. Tressler (Plenum, New York, 1994), p. 253.
Shioya, T. and R. Ishida, “Microscopic fracture modes of brittle polymers in dynamic crack
propagation,” in Dynamic Failure of Materials, edited by H.P. Rossmanith and A.J.
Rosakis (Elsevier Applied Science, Essex, 1991), p. 351.
Shirley, E.L., D.C. Allan, R.M. Martin, and J.D. Joannopoulos, “Extended norm-conserving
pseudopotentials,” Phys. Rev. B 40, 3652 (1989).
Sieradzki, K., G.J. Dienes, A. Paskin, and B. Massoumzadeh, “Atomistics of crack
propagation,” Acta Mettal. 36, 651 (1988).
Sieradzki, K., and R. Li, “Fracture behavior of a solid with random porosity,” Phys. Rev.
Lett. 56, 2509 (1986).
Sieradzki, K., and R.C. Newman, “Brittle behaviour of ductile metals during stress-
corrosion cracking,” Philos. Mag. A 51, 95 (1985).
Sig, G.C., “Some basic problems in fracture mechanics and new concepts,” Eng. Fract.
Mech. 5, 365 (1973).
Sinclair, J.E., and B.R. Lawn, “An atomistic study of cracks in diamond-structure crystals,”
Proc. R. Soc. Lond. A 329, 83 (1972).
Skjeltorp, A.T., and P. Meakin, “Fracture in microsphere monolayers studied by experiment
and computer simulation,” Nature 335, 424 (1988).
Slepyan, L., “Dynamics of a crack in a lattice,” Sov. Phys. Dokl. 26, 538 (1981).
Sloan, S.W., “A fast algorithm for constructing Delaunay triangulations in the plane,” Adv.
Engng. Software, 9, 34 (1987).
Smekal, E., “Zum Bruchvorgang bei sprodem stoffrerhalten unter ein und mehrachsinen
beanspruchungen,” Osterr. Ing. Arch. 7, 49 (1953).
Smith, R.L., S.L. Phoenix, M.R. Greenfield, R.B. Henstenburg, and R.E. Pitt, “Lower-tail
approximations for the probability of failure of three-dimensional fibrous composites
with hexagonal geometry,” Proc. R. Soc. Lond. A 388, 353 (1983).
Smith, W., and T.R. Forester, “Parallel macromolecular simulations and the replicated data
strategy II: The RD-SHAKE algorithm,” Comput. Phys. Commun. 79, 63 (1994).
Smyshlyaev, V.P., and N.A. Fleck, “Bounds and estimates for the overall plastic behavior
of composites with strain gradients effects,” Proc. R. Soc. Lond. A 451, 795 (1995).
Söderberg, M., “Resistive breakdown of inhomogeneous media,” Phys. Rev. B 35, 352
(1987).
Sornette, D., “Weibull-like failure distribution induced by fluctuations in percolation,” J.
Physique 49, 889 (1988).
Sornette, D., “Elasticity and failure of a set of elements loaded in parallel,” J. Phys. A 22,
L243 (1989).
Sornette, D., and C. Vanneste, “Dynamics and memory effects in rupture of thermal fuse
networks,” Phys. Rev. Lett. 68, 612 (1992).
Sornette, D., and C. Vanneste, “Dendrites and fronts in a model of dynamical rupture with
damage,” Phys. Rev. E 50, 4327 (1994).
Soules, T.F., and R.F. Busbey, “The rheological properties and fracture of a molecular
dynamic simulation of sodium silicate glass,” J. Chem. Phys. 78, 6307 (1983).
Spence, J.C.H., Y.M. Huang, and O. Sankey, “Lattice trapping and surface reconstruction
for silicon cleavage on (111). Ab-initio quantum molecular dynamics calculations,” Acta
Mettal. Mater. 41, 2815 (1993).
Spitzig, W.A., R.E. Smelser, and O. Richmond, “The evolution of damage and fracture in
iron compacts with various initial porosities,” Acta Metall. Mater. 36, 1201 (1998).
References 623
Sridhar, N., W. Yang, D.J. Srolovitz, and E.R. Fuller, Jr., “Microstructral mechanics model
of anisotropic-thermal-expansion-induced microcracking,” J. Am. Ceramic Soc. 77, 1123
(1994).
Srolovitz, D.J., and P.D. Beale, “Computer simulation of failure in an elastic model with
randomly distributed defects,” J. Amer. Cer. Soc. 71, 362 (1988).
Stadler, J., R. Mikulla, and H.-R. Trebin, “IMD: a software package for molecular dynamics
studies on parallel computers,” Int. J. Mod. Phys. C 8, 1131 (1997).
Stanley, H.E., and P. Meakin, “Multifractal phenomena in physics and chemistry,” Nature
335, 405 (1988).
Starkloff, T., and J.D. Joannopoulos, “Local pseudopotential theory for transition metals,”
Phys. Rev. B 16, 5212 (1977).
Stauffer, D., and A. Aharony, Introduction to Percolation Theory, 2nd ed. (Taylor and
Francis, London, 1992).
Steele, W.A., “The physical interaction of gases with crystalline solids. I. Gas-solid energies
and properties of isolated adsorbed atoms,” Surf. Sci. 36, 317 (1973).
Stephens, M.D., and M. Sahimi, “Distribution of fracture strengths in disordered continua,”
Phys. Rev. B 36, 8656 (1987).
Stinchcombe, R.B., “Conductivity and spin-wave stiffness in disordered systems-an exactly
soluble model,” J. Phys. C 7, 179 (1974).
Stinchcombe, R.B., P.M. Duxbury, and P. Shukla, “The minimum gap on diluted Cayley
trees,” J. Phys. A 19, 3903 (1986).
Stillinger, F.H., and T.A. Weber, “Computer simulation of local order in condensed phases
of silicon,” Phys. Rev. B 31, 5262 (1985).
St. John, C., “The brittle-to-ductile transition in pre-cleaved silicon single crystals,” Philos.
Mag. 32, 1193 (1975).
Stockman, M.I., “Inhomogeneous eigenmode localization, chaos, and correlations in large
disordered clusters,” Phys. Rev. E 56, 6494 (1997).
Stockman, M.I., L.N. Pandey, and T.F. George, “Inhomogeneous localization of polar
eigenmodes in fractals,” Phys. Rev. B 53, 2183 (1996).
Stockman, M.I., L.N. Pandey, L.S. Muratov, and T.F. George, “Giant fluctuations of local
optical fields in fractal clusters,” Phys. Rev. Lett. 72, 2486 (1994).
Stockman, M.I., L.N. Pandey, L.S. Muratov, and T.F. George, “Optical absorption and
localization of eigenmodes in disordered clusters,” Phys. Rev. B 51 185 (1995)
Stoddard, S.D., and J. Ford, “Numerical experiments on the stochastic behavior of a
Lennard–Jones gas system,” Phys. Rev. A 8, 1504 (1973).
Straley, J.P., “Random resistor tree in an applied field,” J. Phys. C 10, 3009 (1977).
Straley, J.P., and S.W. Kenkel, “Percolation theory for nonlinear conductors,” Phys. Rev. B
29, 6299 (1984).
Strang, G., and G.J. Fix, An Analysis of the Finite Element Method (Prentice-Hall, Englewood
Cliffs, 1973).
Stroh, A.N., “A theory of the fracture of metals,” Adv. Phys. 6, 418 (1957).
Stroud, D., and P.M. Hui, “Nonlinear susceptibilities of granular matter,” Phys. Rev. B 37,
8719 (1988).
Stroud, D., and V.E. Wood, “Decoupling approximation for the nonlinear-optical response
of composite media,” J. Opt. Soc. Am. B6, 778 (1989).
Stroud, D., and X. Zhang, “Cubic nonlinearities in small-particle composites: local-field
induced giant enhancements,” Physica A 207, 55 (1994).
Suquet, P., “Analyse limite et homogénéisation,” C. R. Acad. Sci. Paris Ser. II 296, 1355
(1983).
624 References
Termonia, Y., and P. Meakin, “Formation of fractal cracks in a kinetic fracture model,”
Nature 320, 429 (1986).
Termonia, Y., P. Meakin, and P. Smith, “Theoretical study of the molecular weight on the
maximum tensile strength of polymer fibre,” Macromolecules 18, 2246 (1985).
Termonia, Y., P. Meakin, and P. Smith, “Theoretical study of the influence of strain rate and
temperature on the maximum strength of perfectly ordered and oriented polyethylene,”
Macromolecules 19, 154 (1986).
Termonia, Y., and P. Smith, “Kinetic model for tensile deformation of polymers. 1. Effect
of Molecular Weight,” Macromolecules 20, 835 (1987).
Termonia, Y., and P. Smith, “Kinetic Model for tensile deformation of polymers. 2. Effect
of entanglement spacing,” Macromolecules 21, 2184 (1988).
Tersoff, J., “New empirical approach for the structure and energy of covalent systems,”
Phys. Rev. B 37, 6991 (1988).
Tersoff, J., “Modelling solid-state chemistry: Interatomic potentials for multicomponent
systems,” Phys. Rev. B 39, 5566 (1989).
Theocaris, P.S., and E.E. Gdoutos, “An optical method for determining opening-mode and
edge sliding-mode stress intensity factors,” J. Appl. Mech. 39, 91 (1972).
Theocaris, P.S., and H.G. Georgiadis, “Bifurcation predictions for moving cracks by the
T-criterion,” Int. J. Fract. 29, 181 (1985).
Thijssen, J.M., Computational Physics (Cambridge University Press, Cambridge, 1999).
Thomson, R., “The physics of fracture,” Solid State Phys. 39, 1 (1986).
Thomson, R., C. Hsieh, and V. Rana, “Lattice trapping of fracture cracks,” J. Appl. Phys.
42, 3154 (1971).
Tokatly, I.V., and O. Pankratov, “Many-body diagrammatic expansion in a Kohn–Sham
basis: Implications for time-dependent density functional theory of excited states,” Phys.
Rev. Lett. 86, 2078 (2001).
Torquato, S., “Electrical conductivity of two-phase disordered composite media,” J. Appl.
Phys. 58, 3790 (1985a).
Torquato, S., “Bulk properties of two-phase disordered media. II. Effective conductivity of
a dilute suspension of penetrable spheres,” J. Chem. Phys. 83, 4776 (1985b).
Torrie, G.M., and J.P. Valleau, “Nonphysical sampling distributions in Monte Carlo free-
energy estimation: Umbrella sampling,” J. Comput. Phys. 23, 157 (1977).
Tsuruta, K., A. Omeltchenko, R.K. Kalia, and P. Vashishta, “Early stages of sintering of sil-
icon nitride nanoclusters: a molecular-dynamics study on parallel machines,” Europhys.
Lett. 33, 441 (1996).
Tua, P.F., and J. Bernasconi, “Monte Carlo simulations of two-dimensional randomly diluted
networks of nonlinear resistors,” Phys. Rev. B 37, 1986 (1988).
Tvergaard, V., “Analysis of tensile properties for a whisker-reinforced metal-matrix
composite,” Acta Metall. Mater. 38, 185 (1990).
Tzschichholz, F., “Peeling instability in Cosserat-like media,” Phys. Rev. B 45, 12691
(1992).
Tzschichholz, F., “Fracturing of brittle homogeneous solids: finite-size scalings,” Phys. Rev.
B 52, 9270 (1995).
Tzschichholz, F., and H.J. Herrmann, “Reaction-diffusion model for the hydration and
setting of cement,” Phys. Rev. E 53, 2629 (1996).
Tzschichholz, F., H.J. Herrmann, H.E. Roman, and M. Puff, “Beam model for hydraulic
fracturing,” Phys. Rev. B 49, 7056 (1994).
Uenoyama, T., L. Esaki, and H. Kotera, “Theory of stability in a nonlinear resistive network,”
Appl. Phys. Lett. 61, 363 (1992).
626 References
Underwood, E.E., and K. Banerji, “Fractals in fractography,” Mater. Sci. Eng. 80, 1 (1986).
Van Damme, H., F. Obrecht, P. Levitz, L. Gatineau, and C. Laroche, “Fractal viscous
fingering in clay slurries,” Nature 320, 731 (1986).
Van Damme, H., C. Laroche, and L. Gatineau, “Radial fingering in viscoelastic media, an
experimental study,” Revue Phys. Appl. 22, 241 (1987a).
Van Damme, H., C. Laroche, L. Gatineau, and P. Levitz, “Viscoelastic effects in fingering
between miscible fluids,” J. Physique 48, 1121 (1987b).
van den Born, I.C., A. Santen, H.D. Hoekstra, and J.Th.M. De Hosson, “Mechanical strength
of highly porous ceramics,” Phys. Rev. B 43, 3794 (1991).
Vanderbilt, D., “Soft self-consistent pseudopotentials in a generalized eigenvalue formal-
ism,” Phys. Rev. B 41, 7892 (1990).
Vanneste, C., and D. Sornette, “The dynamical thermal fuse model,” J. Phys. France I 2,
1621 (1992).
Verges, J.A., “Localization length in a random magnetic field,” Phys. Rev. B 57, 870 (1998).
Verlet, L., “"Computer "experiments" on classical fluids. I. Thermodynamical properties of
Lennard–Jones molecules,” Phys. Rev. 159, 98 (1967).
Voigt, W., “Ueber die Bezienhung zwischen den beiden Elasticitäts-constanten isotroper,”
Ann. Physik 38, 573 (1889).
Vold, M.J., “Computer simulation of floc formation in a colloidal suspension,” J. Colloid
Inter. Sci. 18, 684 (1963).
Voss, R.F., “Random fractal forgeries,” in Fundamental Algorithms for Computer Graphics,
edited by R.A. Earnshaw, NATO ASI Series, Vol. 17 (Springer-Verlag, Heidelberg, 1985),
p. 805.
Vu, B.Q., and V.K. Kinra, “Brittle fracture of plates in tension: static field radiated by a
suddenly stopping crack,” Eng. Fract. Mech. 15 107 (1981).
Wagner, N.J., B.L. Holian, and A.F. Voter, “Molecular-dynamics simulations of two-
dimensional materials at high strain rates,” Phys. Rev. A 45, 8457 (1992).
Wan, W.M.V., H.C. Lee, P.M. Hui, and K.W. Yu, “Mean-field theory of strongly nonlinear
random composites: Strong power-law nonlinearity and scaling behavior,” Phys. Rev. B
54, 3946 (1996).
Wang, Z.G., D.L. Chen, X.X. Jiang, S.H. Ai, and C.H. Shih, “Relationship between fractal
dimension and fatigue threshold value in dual-phase steels,” Scripta Metall. 22, 827
(1988).
Wang, Z.-G., U. Landman, R.L. Blumberg Selinger, and W.A. Gelbart, “Molecular-dynamics
study of elasticity and failure of ideal solids,” Phys. Rev. B 44, 378 (1991).
Washabaugh, P.D., and W.G. Knauss, “A reconciliation of dynamic crack velocity and
Rayleigh wave speed in isotropic brittle solids,” Int. J. Fract. 65, 97 (1994).
Webb, III, E.B., and G.S. Grest, “Liquid/vapor surface tension of metals: Embedded atom
method with charge gradient corrections,” Phys. Rev. Lett. 86, 2066 (2001).
Weichert, R., and K. Schonert, “On the temperature at the tip of a fast running crack,” J.
Mech. Phys. Solids 22, 127 (1974).
Weiner, J.H., and M. Pear, “Crack and dislocation propagation in an idealized crystal model,”
J. Appl. Phys. 46, 2398 (1975).
Whitehead, S., Dielectric Breakdown of Solids (Clarendon, Oxford, 1951).
Wiener, O., “Die theorie des mischk´’orpers fr̈ das feld des stationären strömung,” Math.-
Physichen Klasse der Königl. Sächsischen Gesellschaft der Wissenschaften 32, 509
(1912).
Wiesmann, H.J., and H.R. Zeller, “A fractal model of dielectric breakdown and
prebreakdown in solid dielectrics,” J. Appl. Phys. 60, 1770 (1986).
References 627
Williford, R.E., “Scaling similarities between fracture surfaces, energies, and a structure
parameter,” Scripta Metall. 22, 197 (1988).
Willis, J.R., “Bounds and self-consistent estimates for the overall moduli of anisotropic
composites,” J. Mech. Phys. Solids 25, 185 (1977).
Willis, J.R., “Variational principles and bounds for the overall properties of composites,”
in Continuum Models and Discrete Systems (CMDS 2), edited by J. Provan (University
of Waterloo Press, Waterloo, Canada, 1978), p. 185.
Willis, J.R., “Variational and related methods for the overall properties of composites,” Adv.
Appl. Mech. 21, l (1981).
Willis, J.R., “The overall response of composite materials,” ASME J. Appl. Mech. 50, 1202
(1983).
Willis, J.R., “Variational estimates for the overall response of an inhomogeneous nonlinear
dielectric,” in Homogenization and Effective Moduli of Materials and Media, edited by
J.L. Ericksen, D. Kinderlehrer, R. Kohn, and J.-L. Lions (Springer, New York, 1986),
p. 247.
Willis, J.R., “The structure of overall constitutive relations for a class of nonlinear
composites,” IMA J. Appl. Math. 43, 231 (1989a).
Willis, J.R., in Micromechanics and Inhomogeniety, the Toshio Mura Anniversary Volume,
edited by G.J. Weng, M. Taya, and H. Abe (Springer, New York, 1989b), p. 581.
Willis, J.R., Elasticity: Mathematical Methods and Applications (Halston Press, New York,
1990), p. 397.
Willis, J.R., “On methods for bounding the overall properties of nonlinear composites,”
Phys. Solids 39, 73 (1991).
Willis, J.R., “On method for bounding the overall properties of nonlinear composites:
Correction and addition,” J. Mech. Phys. Solids 40, 441 (1992).
Willis, J.R., and A.B. Movchan, “Dynamic weight functions for a moving crack. I. Mode I
loading,” J. Mech. Phys. Solids 43, 319 (1995).
Willis, J.R., and A.B. Movchan, “Three dimensional dynamic perturbation of a propagating
crack,” J. Mech. Phys. Solids 45, 591 (1997).
Winkler, S., D.A. Shockey, and D.R. Curran, “Crack propagation at supersonic velocities.
I,” Int. J. Fract. 6, 151 (1970).
Witten, T.A., and L.M. Sander, “Diffusion-limited aggregation, a kinetic critical phe-
nomenon,” Phys. Rev. Lett. 47, 1400 (1981).
Wood, W.W., and F.R. Parker, “Monte Carlo equation of state of molecules interacting
with the Lennard–Jones Potential. I. A supercritical isotherm at about twice the critical
temperature,” J. Chem. Phys. 27, 720 (1957).
Wright, D.C., D.J. Bergman, and Y. Kantor, “Resistance fluctuations in random resistor
networks above and below the percolation threshold,” Phys. Rev. B 33, 396 (1986).
Wu, B.Q., and P.L. Leath, “Failure probabilities and tough-brittle crossover of heteroge-
neous materials with continuous disorder,” Phys. Rev. B 59, 4002 (1999).
Wu, B.Q., and P.L. Leath, “Fracture strength of one-dimensional systems with continuous
disorder: A single-crack approximation,” Phys. Rev. B 61, 15028 (2000).
Wu, K., and R.M. Bradley, “Theory of electromigration failure in polycrystalline metal
films,” Phys. Rev. B 50, 12468 (1994).
Xu, G., A.S. Argon, and M. Ortiz, “Nucleation of dislocation from crack tips under mixed
modes of loading: Implications for brittle against ductile behaviour of crystals,” Philos.
Mag. 72, 415 (1995).
628 References
Xu, L., M. Sahimi, and T.T. Tsotsis, “Nonequilibrium molecular dynamics simulations of
transport and separation of gas mixtures in nanoporous materials,” Phys. Rev. E 62, 6942
(2000b).
Xu, L., M.G. Sedigh, M. Sahimi, and T.T. Tsotsis, “Nonequilibrium molecular dynamics
simulation of transport of gas mixtures in nanopores,” Phys. Rev. Lett. 80, 3511 (1998).
Xu, L., T.T. Tsotsis, and M. Sahimi, “Nonequilibrium molecular dynamics simulation of
transport and separation of gases in carbon nanopores. I. Basic results,” J. Chem. Phys.
111, 3252 (1999).
Xu, L., M.G. Sedigh, T.T. Tsotsis, and M. Sahimi, “Nonequilibrium molecular dynamics
simulation of transport and separation of gases in carbon nanopores. II. Binary and ternary
mixtures and comparison with the experimental data,” J. Chem. Phys. 112, 910 (2000a).
Xu, X.P., and A. Needleman, “Numerical simulations of fast crack growth in brittle solids,”
J. Mech. Phys. Solids 42 1397(1994).
Yagil, Y., G. Deutscher, and D.J. Bergman, “Electrical breakdown measurements of
semicontinuous metal films,” Phys. Rev. Lett. 69, 1423 (1992).
Yagil, Y., G. Deutscher, and D.J. Bergman, “The role of microgeometry in the electrical
breakdown of metal-insulator mixtures,” Int. J. Mod. Phys. B 7, 3353 (1993).
Yagil, Y., G. Deutscher, and D.J. Bergman, “Nonlinear electrical response and breakdown
of semicontinuous metal films,” Physica A 207, 323 (1994).
Yagil, Y., P. Gadanne, C. Julien, and G. Deutscher, “Optical properties of thin semicontinuous
gold films over a wavelength range of 2.5 to 500 µm,” Phys. Rev. B 46, 2503 (1992).
Yan, Y., G. Li, and L.M. Sander, “Fracture growth in 2d elastic networks with Born model,”
Europhys. Lett. 10, 7 (1989).
Yang, C.S., and P.M. Hui, “Effective nonlinear response in random nonlinear resistor
networks: numerical studies,” Phys. Rev. B 44, 12559 (1991).
Yoffe, E.H., “The moving Griffith crack,” Philos. Mag. (series 7) 42, 739 (1951).
Yonenaga, I., and K. Sumino, “Mechanical properties and dislocation dynamics of GaP,” J.
Mater. Res. 4, 355 (1989).
Young, A.P., “Melting and the vector Coulomb gas in two dimensions,” Phys. Rev. B 19,
1855 (1979).
Yoon, Y.-G., M.S.C. Mazzoni, H.J. Choi, J. Ihm, and S.G. Louie, “Structural deformation
and intertube conductance of crossed carbon nanotube junctions,” Phys. Rev. Lett. 86,
688 (2001).
Yu, K.W., and G.Q. Gu, “Electrostatic boundary-value problems of nonlinear media: a
perturbation approach,” Phys. Lett. A 168, 313 (1992).
Yu, K.W., and G.Q. Gu, “Effective conductivity of nonlinear composites. II. Effective-
medium approximation,” Phys. Rev. B 47, 7568 (1993).
Yu, K.W., and G.Q. Gu, “Variational calculation of strongly nonlinear composites,” Phys.
Lett. A 193, 311 (1994).
Yu, K.W., and G.Q. Gu, “Effective conductivity of strongly nonlinear composites:
Variational approach,” Phys. Lett. A 205, 295 (1995).
Yu, K.W., and P.M. Hui, “Percolation effects in two-component nonlinear composites:
Crossover from linear to nonlinear behavior,” Phys. Rev. B 50, 13327 (1994).
Yu, K.W., P.M. Hui, and D. Stroud, “Effective dielectric response of nonlinear composites,”
Phys. Rev. B 47, 14150 (1993).
Yuse, A., and M. Sano, “Transition between crack patterns in quenched glass plates,” Nature
362 329 (1993).
Zapperi, S., P. Ray, H.E. Stanley, and A. Vespignani, “First-order transition in the breakdown
of disordered media,” Phys. Rev. Lett. 78, 1408 (1997).
References 629
Zehnder, A.T., and A.J. Rosakis, “On the temperature distribution at the vicinity of
dynamically propagating cracks in 4340 steel,” J. Mech. Phys. Solids 29, 385 (1991).
Zeng, X.C., D.J. Bergman, P.M. Hui, and D. Stroud, “Effective-medium theory for weakly
nonlinear composites,” Phys. Rev. B 38, 10970 (1988).
Zeng, X.C., P.M. Hui, D.J. Bergman, and D. Stroud, “Mean field theory for weakly nonlinear
composites,” Physica A 157, 192 (1989).
Zepeda-Ruiz, L.A., D. Maroudas, and W.H. Weinberg, “Theoretical study of the energet-
ics, strain fields, and semicoherent interface structures in layer-by-layer semiconductor
heteroepitaxy,” J. Appl. Phys. 85, 3677 (1999).
Zhang, G.M., “Cross-over components in percolation superconductor-nonlinear-conductor
mixtures,” Phys. Rev. B 53, 20 (1996a).
Zhang, G.M., “Higher order nonlinear response in random resistor networks: numerical
studies for arbitrary nonlinearity,” Z. Phys. B 99, 599 (1996b).
Zhang, X., and D. Stroud, “Numerical studies of the nonlinear properties of composites,”
Phys. Rev. B 49, 944 (1994).
Zhang, Y.M., and T.C. Wang, “Lattice Instability at a fast moving crack tip,” J. Appl. Phys.
80, 4332 (1996).
Zhenyi, M., S.C. Langford, J.T. Dickinson, M.H. Eengelhard, and D.R. Baer, “Scanning
tunneling microscope observations of MgO fracture surfaces,” J. Mater. Res. 6, 183
(1990).
Zhou, S.J., D.M. Beazley, P.S. Lomdahl, and B.L. Holian, “Large-scale molecular dynamics
simulations of three-dimensional ductile failure,” Phys. Rev. Lett. 78, 479 (1997).
Zhou, S.J., A.E. Carlsson, and R. Thomson, “Dislocation nucleation and crack stability:
lattice Green’s-function treatment of cracks in a model hexagonal lattice,” Phys. Rev. B
47, 7710 (1993).
Zhou, S.J., A.E. Carlsson, and R. Thomson, “Crack blunting effects on dislocation emission
from cracks,” Phys. Rev. Lett. 72, 852 (1994).
Zhou, S.J., and W.A. Curtin, “Failure of fiber composites: a lattice Green function model,”
Acta Metall. Matter 43, 3093 (1995).
Zhou, S.J., P.S. Lomdahl, R. Thomson, B.L. Holian, “Dynamic crack processes via
molecular dynamics,” Phys. Rev. Lett. 76, 2318 (1996).
Zielinski, W., H. Huang, S. Venkataraman, and W.W. Gerberich, “Dislocation distribution
under a microindentation into an iron-silicon single crystal,” Philos. Mag. A 72, 1221
(1995).
Zimmerman, C., W. Klemm, and K. Schonert, “Dynamic energy release rate and fracture
heat in polymethylmethacrylate (PMMA) and a high strength steel,” Eng. Fract. Mech.
20, 777 (1984).
Index