Sei sulla pagina 1di 358

Random Materials and Processes

Series Editors:
H.Eugene Stanley and Etienne Guyon

Other Volumes in this series:

Hydrodynamics of Dispersed Media


J.P. Hulin, A.M. Cazabat,
E. Guyon, F. Carmona (editors)
ISBN 0444 883568

NORTH-HOLLAND
AMSTERDAM · OXFORD · NEW YORK · TOKYO
Statistical Models for the
Fracture of Disordered Media

Edited by:

Hans J. Herrmann
S.Ph.T.
CENSaclay
Gif-sur-Yvette, France

Stephane Roux
LPMMH
ESPCI
Paris, France

1990
NORTH-HOLLAND
AMSTERDAM · OXFORD · NEW YORK · TOKYO
North-Holland
Elsevier Science Publishers B.V.
Sara Burgerhartstraat 25
P.O. Box 211
1000 AE Amsterdam
The Netherlands

Sole distributors for the U.S.A. and Canada:


Elsevier Science Publishing Company, Inc.
655 Avenue of the Americas
New York, N.Y. 10010
U.S.A.

Cover illustration:
A single crack growing in a dielectric medium with a quenched disorder in the local fracture thres-
holds. The lattice is a 80 x 80 triangular lattice, and periodic boundary conditions are implemented.

Library of Congress Cataloging-in-Publication Data

Statistical models for the fracture of disordered media / edited by


Hans J. Herrmann, Stephane Roux.
p. cm. -- (Random materials and processes)
Includes bibliographical references.
ISBN 0-444-88551 -X - ISBN 0-444-88550-1 (pbk.)
1. Fracture mechanics-Statistical methods. 2. Statistical physics.
I. Herrmann, Hans J. I I . Roux, Stephane. III. Series.
TA409.S73 1990
620. 1'126-dc20 90-6966
CIP

ISBN: 0 444 88551 x (hardbound)


0 444 88550 1 (paperback)

© ELSEVIER SCIENCE PUBLISHERS B.V., 1990

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or
otherwise, without the prior written permission of the Publisher, Elsevier Science Publishers B. V.,
P.O. Box 211, 1000 AE Amsterdam, The Netherlands.

Special regulations for readers in the U.S.A.: This publication has been registered with the
Copyright Clearance Center Inc. (CCC), Salem, Massachusetts. Information can be obtained from
the CCC about conditions under which photocopies of parts of this publication may be made in the
U.S.A. All other copyright questions, including photocopying outside of the U.S.A., should be refer-
red to the Publisher, unless otherwise specified.

No responsibility is assumed by the Publisher for any injury and/or damage to persons or property
as a matter of products liability, negligence or otherwise, or from any use or operation of any
methods, products, instructions or ideas contained in the material herein.

Printed in The Netherlands


v

It is widely believed that between now and the turn of the century a substantial
fraction of basic and applied scientists will be working on problems concerning
structures characterized by the lack of conventional geometrical order. Random
materials constitute a very large part of our natural environment as well as a sub-
stantial fraction of man-made objects. Glasses, polymers and amorphous materials
are among the vast array of examples. On a larger scale, porous media, composites
and suspensions can be mentioned.

Although the basic new ideas are finding widespread application, the field is trou-
bled by the fact that a single coherent set of books that cover these new develop-
ments does not exist. Indeed, since the field is truly interdisciplinary, the written
accounts are as likely to be found in physics and chemistry books and journals as
they are to be found on the shelves labeled "materials science" or "engineering".

The basic ideas in the subject (such as percolation, fractals, random fields, . . . )
were generally introduced in connection with microscopic problems. It has, how-
ever, become clear to a community of scientists that these ideas are relevant to a
broad spectrum of applications dealing with materials science where the disorder
is at an intermediate or "rnesoscopic" scale. Not only static structures but also
dynamic effects should be included. Examples are found in the growth processes
of aggregates, in chemical dispersion within a fractured rock, in oceans, or in the
structure of a flame front. In order to encompass this whole field, we are intro-
ducing this coherent series of books designed to reach engineers, scientists and
mathematicians interested in recent progress in this multidisciplinary field. Each
volume strives to be self-contained and comprehensible to readers from outside the
field

The Series Editors


vii

Preface

When Stephane Roux and Hans Herrmann asked me to write this introduction,
they knew my reluctant attitude towards their studies: in the past, I had found it
hard, and possibly dangerous, to build theories of fracture in a random medium
at a time where we did not really understand the motion of a fracture line in
homogeneous media, such as a polymer matrix. But Roux and Herrmann were
confident - and they were right. Their book does answer most of my queries. For
instance, I have often complained about theorists taking concepts from one field
(say F) and bringing them into another field (G) without having really digested
the experimental knowledge of ((?): but here the book starts from facts - covering
metals, ceramics, polymers, cements, together with the remarkable example of
muds (with variable clay content).
I have also little enthusiasm for "books" which are a loose aggregate of
individual articles - a common plague of our time, resulting from the united forces
of commercial editors and of funding agencies. But in the present text, I perceive
a real effort towards unification.
The crucial part of the book describes, and systemizes, numerical experi-
ments where a random medium, under increasing tensions, nucleates microcracks,
which then grow and interact, up to a moment of macroscopic rupture. The sim-
ulations do reveal certain scaling laws (up to a time of total rupture), and this is
an important advance.
Is this behavior, specific of strongly disordered solids under quasistatic
conditions, really relevant in practice? It will certainly be important for cements
and possibly for seismology; it may also define a useful limit for other, disordered
materials. Let us hope that these challenging questions will provide a meeting point
between mechanics and physics; and let us thank the two coordinators who helped
to build up a common language.

Pierre-Gilles de Gennes
November 1989
IX

Foreword

The breaking of glass, the cracking of concrete walls or the rupture of a wire are
such everyday events that, not surprisingly, fracture has been studied already by
Galileo.M Since the beginning of the century the technological desire to master
the fracture of metals, concrete or polymers has boosted research and has left
behind an overwhelming amount of literature which in some cases has even been
encyclopedically reviewedJ2^ Consequently it seems difficult to say anything simple
and new in this field.
The present book nevertheless tries to take up with this challenge. The
approach to fracture that we want to describe was not conceived by mechanical
engineers or material scientists but is essentially the byproduct of exciting devel-
opments that occurred in the last ten or fifteen years within a branch of theoretical
physics, called statistical physics. We will therefore introduce concepts like "per-
colation" or "fractals" and discuss properties of fracture that are not very often
considered by engineers like the shape of cracks or the local distribution of strain.
A particular aim of this volume is to emphasize the fundamental role disorder plays
in the breaking process.
The merging of concepts from different areas into a coherent new picture
can be very fruitful provided a common language between the various scientific
communities can be found. Therefore the main scope of this volume is pedagogical.
It should be at the same time an overview of fracture mechanics for physicists
and an introduction to new concepts of statistical physics for mechanicians and
engineers. For this reason the first half of this book consists of introductory chapters
and the results that came out of the new approach will be presented in the second
half.
The first chapter gives a general, brief introduction to the subject fracture
reviewing some of the experimental facts and sketching the existing phenomenol-
ogy and the traditional theoretical approaches. The second chapter consists of six
contributions written by insiders who do modern experimental research on differ-
ent materials. This chapter may help to close the gap between the real world and
the theoretical considerations that follow. Chapter 3 exposes the classic formalisms
needed to describe the mechanical response of a system, like elasticity or plasticity.
Chapter 4 presents percolation, the simplest model that describes the essence of
disorder and fractality. Chapter 5 gives a general overview of the various statistical
X Foreword

modelizations of fracture and the approximations involved in each of these models.


Two of these modelizations are presented in the two chapters that follow. New phe-
nomena like multifractaJity and finite size scaling will be encountered. Chapter 8
displays the analogy between fracture and Laplacian growth as found in viscous
fingering or dielectric breakdown. Spectacular simulations of the behaviour of dry-
ing paint are presented in chapter 9. The last chapter contains recent progress in
a field full of interesting open questions: fragmentation.

References
1. G. Galilei, Discorsi e Dimostrazioni matematiche, Leiden, 1638
2. See e.g. H. Liebowitz (ed.), fracture, Vols.I-VII, (Academic Press, New York,
1986)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
©Elsevier Science Publishers B.V. (North-Holland), 1990 1

1. Introduction to basic notions and facts

Hans J. Herrmann*

This first chapter is particularly simple. The main goal is to give physicists
an idea about the phenomenon of fracture without entering into many details. I
hope the experts will allow for the many oversimplifications, omissions and the
lack of precision. The reader who wants to know more can read one of the many
textbooks^ that have been written on the subject or consult the classical collection
of review articles edited by LiebowitzJ2^

1.1 Objectives
Every solid breaks if a sufficiently large load is applied on it. The value
of this load as well as the shape and other characteristics of the resulting crack
strongly depend on the material and on the way how the load has been applied.
To illustrate how different situations can arise we show in figure 1.1 four different
shapes of cracks obtained under various conditions and for different materials.
Fig. 1.1a shows the crack of an alloy that started on a free surface obtained under
an uniaxial tension in a corroding environment. Fig. 1.1b shows various stages of
the breaking of a thin sheet of clay under shear. In fig. 1.1c a drop of acid was
placed on the surface of a plexiglass plate some time ago and fig. l.ld shows the
surface of a drying block of concrete.
In order to understand why and how a given sample breaks one can pursue
various approaches depending essentially on the length scale in which one is inter-
ested. As depicted in figure 1.2 several disciplines are involved in this enterprise.
Ranging from atomic scales to the size of grains the mechanisms involved in frac-
ture are very diverse and strongly vary from one material to the other. Generically
these microscopic scales are studied by material scientists, in fact, there are at least
three different families of materials: metals, rocks and polymers each studied by
its own community. On human scales, i.e. from one centimeter upwards fracture is
a problem of civil and mechanical engineering. In this range the aim is the design
of structures such as to prevent their failure and this is achieved by relying heav-
ily on experience. In the intermediate range, i.e. between microns and centimeters
t Service de Physique Theorique, CEN Saclay, F-91191 Gif sur Yvette Cedex, France.
2 H. J. Herrmann / Introduction to basic notions and facts

Figure 1.1 (a) Intergranular stress corrosion crack of Inconel 600 in high temperature water,
taken from ref. [3]. (b) Four stages of the development of a fault network on the surface of a
sample of wet clay under shear, by Z. Reches from ref. [4]. (c) Surface of a plexiglass plate after
the reaction of a drop of acid by G. Ananthakrishna, Kalpakkam, India, (d) Surface of a 5 cm
thick plate of drying concrete by P. Acker, (LCPC) and L. Oger, (ESPCI), Paris.

FRACTURE FRACTURE PROCESSES PLASTICITY TESTING APPLICATIONS


and CRITERIA

£ ^ilhl^ii,

JL
10,-6
9
10' 10" 10"' IIP 10 10 meters
materials science engineering
applied mechanics
h« — *Ί
fracture mechanics

Figure 1.2 Schematic representation of the study of fracture on different lengths scales.

the mechanical behaviour can be described rather well by continuum equations in


which the material is just characterized by some parameters and where only few
types of behaviour, like "elastic", "plastic" or "viscoelastic", are distinguished. In
this book we will mainly focus on this intermediate regime and most of the concepts
that will be introduced will apply to this regime.
The fracture of an ideal crystal can be handled theoretically and various
properties can be calculated from first principles. Unfortunately, however, the force
needed to break the crystal that one obtains from these calculations is already
2. The mechanical response of a soHd 3

several orders of magnitude larger than forces measured experimentally on real


materials. The reason for this discrepancy is that real substances have "disorder",
i.e. they have deviations from the perfect crystal structure and it happens that the
process of fracture is extremely sensitive to disorder. Why this is so and which are
the macroscopic effects of disorder on fracture is one of the central issues of this
book.
Microscopically disorder can mean many things: vacancies, inclusions, com-
posite structure, dislocations or even grain boundaries but on the length scales we
are interested in, these spatial inhomogeneities can be reduced to a random noise
in the material properties or, on the computer, to sequences of (quasi-)random
numbers. Cast into this form randomness is familiar to statistical physicists. It is
known that in such a formalism one can find critical phenomena, like "percola-
tion", one can see "fractals", specially if one has an unstable growth phenomenon
and one can encounter unusual statistics, like "non-self-averaging" quantities and
"multifractal" probability distributions. How concretely all this appears in fracture
that is the content of this volume. Chapter 4 is devoted to introduce the notions
of disorder.

1.2 The mechanical response of a solid


Before discussing the breaking of a solid it is useful to understand the me-
chanical behaviour of a flawless, homogeneous medium. On length scales on which
one does not feel any more the chemistry or the texture of the material its mechan-
ical behaviour is determined by a few quantities like the "toughness", the "yield
point" or the "elastic moduli" which enter in what are called the "constitutive
relations" or "rheological laws". These relations tell how the material reacts to a
local application of a force. A well-known example for such a relation is Hooke's
law. Constitutive relations are directly obtained from the experiment.
Let us consider the classical experiment by Young. A homogeneous bar of
length L and cross section w x w is pulled in the direction of the length with a force
F , i.e. submitted to a uniaxial tension. Due to the load the block is elongated by
AL and its width decreased by Aw (see also figure 1.2 of chapter 3). The relative
elongation δ = AL/L can be measured as a function of F and in fig. 1.3 we see
three typical situations.
For small elongations the relation between F and 6 is usually linear and
reversible and one can apply the formalism of linear elasticity which will be dis-
cussed in chapter 3. An isotropic medium is characterized in this linear theory by
just two elastic moduli, the "Young modulus" E and the "Poisson ratio" v defined
by Hooke's law σ = F/w2 = Εδ = -(E/v)(Aw/w) where E > 0 and - 1 < v < \
for thermodynamical reasons; σ is called the "stress". For simplicity most of the
fracture models in this book will be considered in this linear approximation only.
If the force is increased beyond a certain material-dependent value Fs one
sees deviations from the linear law. In some specific cases it can even happen that
4 HJ. Herrmann / Introduction to basic notions and facts

δ o —»-
6
(a) (b) (c)

Figure 1.3 Constitutive relations obtained through Young's experiment for (a) copper, (b) mild
steel and (c) the ideal case of perfect plasticity.

the relation is non-linear already for infinitesimal forces due to force-dependent


internal contacts (F oc δ3^2 is Hertz's law for spheres^ and F oc £ 3 " 4 is found for
a random packing of cylinders^). In these cases the elastic moduli are generalized
to be force-dependent functions and one calls k = dF/dAL the "stiffness" and its
inverse the "compliance" of the system. Numerically these cases can be handled
by inserting the experimental, non-linear constitutive relation into the relaxation
algorithms that one also uses to solve the linear equations.
Up to now we only encountered "elastic" or reversible response, i.e. that
the system goes back to exactly the original shape when the force is reset to
zero. If a certain material-dependent force F Y , called the "yield point", is passed
this reversibility is lost and one finds "plastic" behaviour. Beyond i*V, let's say
at A in fig. 1.3a, a finite, permanent elongation, called "plastic deformation" or
"dilatancy", remains when the force is reset to zero which is the point B in fig. 1.3a.
Plasticity is due to flow inside the material, commonly flow along crystal
planes. In the ideal case of "perfect plasticity", i.e. when dF/dAL = 0 beyond FY
as shown in fig. 1.3c, the flow is indistinguishable from that of some fluids. In fact
there are "non-Newtonian" fluids^ like "Bingham fluids", which need a threshold
pressure (more precisely shear stress) to flow, analogous to the yield point so that
for these particular cases the distinction between fluid and solid becomes rather
semantic.
In the majority of cases, however, one has a finite "strain-hardening",
i.e. dF/dAL is non-zero beyond Fy as is the case in figs. 1.3a and 1.3b. Strain-
hardening implies that once Fy is passed and the force is reset to zero and increased
again the system becomes stifler as can be seen from fig. 1.3a: The curve BA is
steeper than OA. The internal flow has modified the material and made it harder,
a process of technological importance. In some cases as the one shown in fig. 1.3c
an extra force is needed to unleash plastic flow, in fact inside the material, bands
of dislocations ("Lüders bands") are set in motion at this point. The resulting
instability is, however, rapidly controlled by the strain hardening.
3. Phenomenology of fracture 5

The mathematical formalism needed to describe plasticity has to include


the dependence on the history of force applications and the fact that for any finite
dilatancy the system is elastic as long as the applied force does not exceed FY.
This non-linear formalism which in its present form was only developed in the last
decades will be described with more detail in chapter 3 but since it is rather heavy
it has not yet been implemented into the statistical fracture models that this book
is concerned about. Plasticity is, however, a very common phenomenon and the
generalization of the approaches presented in this book to the plastic case is an
important task for the future.
The deformation of a solid can cause energy dissipation, e.g. heating, and
as a consequence the response of the system will not be immediate but delayed
(see also figure 1.3 of chapter 3). This effect, which is called "viscoelasticity" or
"viscoplasticity" depending on if one is below or above the yield point, introduces
at least one characteristic delay time r into the problem. Introducing r-dependent
material constants one can take into account these effects so that for instance in
Hooke's law the force depends on elongations made over a time interval (σ(£) =
/o°° C(r)6(t — r)dr where C(r) is a material-dependent delay kernel). Applying
periodic forces (or imposing periodic elongations) with a frequency ω allows to
measure the dissipation. In the linear case the frequency-dependent formalism is
completely analogous to the case of highly viscous fluids^ and in fact for ωτ <C 1
these fluids behave like elastic solids (see e.g. the clay discussed in chapter 2).
Again we are on the borderline between fluid and solid.

1.3 Phenomenology of fracture


1.3.1 Brittle and ductile
When one reaches a certain force Fc while performing Young's experiment
the sample will break apart. If this happens before getting to the yield point Fc <
FY the system behaves elastic till the breaking and the fracture is called "bnttle".
In the opposite case Fc > Fy the sample breaks in the plastic regime and the
fracture is called "ductile". Typically a fresh potatoe chip breaks brittle and a
chewing gum ductile (at room temperature). Brittle or ductile fracture are not just
properties of the material but depend also on temperature or pressure. In 1911 von
Kärman discovered that Carrara marble which is usually brittle becomes ductile
when submitted to hydrostatic pressure^ (see fig. 1.4b). The force at which that
happens is about three times larger under tension than under compression. Brittle
to ductile transitions were equally found increasing the temperature or decreasing
the speed of the deformation, or through the presence of corroding agents.
It is again useful to consider the dependence between force F and elonga-
tion δ which in the case of fracture is also called the "breaking characteristics".
In fig. 1.4a we see the transition from brittle to ductile as obtained from the ex-
periment on Carrara marble by showing the characteristics at different hydrostatic
pressures σχ. For σχ = 0 and σ\ — 10 MPa one has brittle fracture and the charac-
6 HJ. Herrmann / Introduction to basic notions and facts

a^lOOMPa
L· \l** 1
σι
0^
F c (0)
0\

tl·*
(a) lb)
Figure 1.4 (a) Breaking characteristics for marble under compression for different hydrostatic
pressures σ\. The uniaxial excess force F is defined in (b) where A is the top (or bottom) surface
of the sample on which the force is applied.

teristics terminates rather abruptly at rather small elongations. Therefore techno-


logically speaking brittle fracture is dangerous. At high pressures of σχ = 100 MPa
fracture is ductile and it occurs after the sample has been substantially elongated.
1 MPa = lmegapascal = 106 N/m 2 is in this field the common MKSA unit used
for elastic moduli, pressure and strength of materials. Table 1.1 shows how it is
related to other units that appear in the literature.
In both cases, brittle and ductile, the characteristics follows for small forces
the constitutive relation of the material. For larger forces the material looses stiff-
ness due to changes that occur inside the material before fracture. At a force Fc the
curve reaches its maximum and after that the sample elongates and breaks with-
out any increase in the externally applied force F. The regime after the maximum,
called "postfailure" regime or also sometimes called "catastrophic" regime in this
book, can be obtained experimentally by imposing the elongation of the sample
instead of imposing the force. Fc is the technologically important "breaking force11.

1 kbar 100 MPa


1 dyn/cm 2 0.1 Pa
1 kp/mm 2 981 MPa
1 psi 6894.5 Pa
Table 1.1 Units of pressure (psi = pounds per square inch).

In the case of brittle fracture the characteristics in the postfailure regime


can show very strong variations from sample to sample and for a given sample the
curve is seldom smooth. One distinguishes postfailure behaviour of class I and of
3. Phenomenology of fracture 7

Figure 1.5 Brittle breaking characteristics which in the postfailure regime is of class I (full line)
and of class II (dashed line).

class II depending if the characteristic bends back or not^10,11! as seen in fig. 1.5.
Tennessee marble is typically of class I and basalt of class II. In order to measure
the characteristics for class II one needs a very stiff testing machine and has to
continuously and very quickly adapt force and elongation via a servo-controlJ12^
Brittle fracture is the main subject of this book. It is therefore our aim
to understand why the characteristics has the given shape and how it depends on
the sample size, the disorder or the mechanical material constants. Experimental
facts concerning the brittle breaking characteristics are scattered throughout the
engineering literature. Let us just mention here that for uniaxial compression of bar-
shaped rocks the postfailure stiffness roughly increases linearly with the diameter
and with the length of the barJ 10 !
In ductile fracture very different phenomena appear, the most spectacular
being geometrical instabilities. Let us consider again Young's experiment where
under uniaxial tension an instability called "necking" occurs: After passing the
yield point flow occurs on certain crystal planes rendering the bar slightly thinner
at some regions as compared to the rest ("localization") as sketched in fig. 1.6a.
Since in these regions the cross section is smaller the force per area is larger and
more elongation, i.e. more flow of crystal planes will take place. Consequently the
region will become thinner and thinner and the force per area will finally diverge
so that the sample tears apart at this neck. This is what happens when we stretch
a chewing gum. Another well-known instability, called "buckling", occurs under
uniaxial compression: The axial symmetry is broken and the bar bulges into one
direction as seen in fig. 1.6b. Buckling is the main source of failure of cellular
structures like bee-hivesJ13^ The geometrical instabilities of plasticity are a subject
of much current i n t e r e s t ^ on one hand as a mathematical problem ("Hadamard
instabilities") and on the other hand because of the variety of phenomena they can
generate. However they are beyond the scope of this book.
8 HJ. Herrmann / Introduction to basic notions and facts

t »

\ t
(a) (b)

Figure 1.6 Geometrical instabilities in ductile fracture: (a) necking, (b) buckling.

Figure 1.7 The three modes of loading.

1.3.2 The three-dimensional nature


For the sake of simplicity we have only considered up to now uniaxial ex-
perimental setups. In real applications one has, however, seldom an axial symmetry
of the load, usually both the externally applied forces and the resulting deforma-
tions of the solid are complicated three-dimensional vector fields. To take this into
account the mechanical response of the solid must now be described by a tensorial
formalism. The force is replaced by the "stress tensor" and the elongation by the
"strain tensor", chapter 3 will deal with this in much detail. Conceptually, however,
not much is changed as compared to the uniaxial case.
The three-dimensional nature of the mechanical loads affects of course also
fracture. On a crack three modes of loading can be distinguished as shown in fig. 1.7.
Any arbitrary load can be expressed in a linear theory as a superposition of these
three modes. Details about this can be found in any textbook on fractureJ15^
The breaking force and the characteristics depend of course also on the type
of external load. Besides uniaxial tension and compression we will encounter in this
book cases of shear, uniform dilatation and radial tension with angular symmetry.
3. Phenomenology of fracture 9

SeU _^rrr

NT
Figure 1.8 Amplitude S of the periodic load as a function of the failure life N{.

Since breaking is essentially due to the growth of a crack at its tip it is possible to
reduce the breaking force Fc to a quantity that describes the stress that is needed
at the tip to open the crack further. This material-dependent quantity Kc, which
is called "toughness", will be defined in the next section.
1.3.3 Dynamical effects
Important in real applications are also time-dependent loads. A very com-
mon phenomenon is "fatigue" :^ A body is submitted to the same load many times
and nothing particularly seems to happen; suddenly when the load is applied for
the iVth time it breaks. If the load is periodic in time, e.g. sinusoidal, one calls this
phenomenon "cyclic fatigue" and the number Nf of cycles for failure to occur is also
called the "failure life". The dependence of JVf on the amplitude S of the periodic
load is schematically shown in fig. 1.8 for the case of metals. We see that if the
amplitude is below the "endurance limit" Se the sample will not break at all. An
empirical relation, also called "Besquin relation", of the form N{ oc (S — Se)~a has
been found to hold for exponents a « 8-20. Except for frequencies above 1000 Hz,
frequency has a negligible effect on fatigue in metals. The curve of fig. 1.8 and
therefore also Se depend not only on the material but also on the type of load. For
uniaxial tension the endurance limit is in general much larger than for shear.
Shocks are another common case of time-dependent loads. They can for
instance be generated by hitting the sample with a hammer and are also encoun-
tered in explosions. Usually at the shock the energy that is locally furnished exceeds
many times the energy necessary for fracture. Partly this energy goes into the ki-
netic energy of the propagating cracks, which therefore propagate very fast. The
excess energy produces crack-branching and fragmentation, both phenomena of
high current interest which will come up again in this book.
Another important time-dependent effect of fracture is "stress corrosion
10 H.J. Herrmann / Introduction to basic notions and facts

cracking". It was discovered^ that the toughness of glass is decreased by 20% if


it is put into a humid environment. The reason is that water penetrates into the
glass at the crack tip because there the crystalline structure is more open. Inside
the glass the water forms a base with existing sodium ions and this basic liquid
corrodes the crack tip region lowering its toughness and enhancing the possibility
for brittle fracture. It turned out that stress corrosion cracking is in most cases the
responsible mechanism for the tearing of the aluminium or titanium alloys that are
used to build airplanes and is therefore an important source of accidents. For this
reason industrial and military agencies have made an important effort to better
understand the phenomenonM7^ This is also the reason why it is forbidden to take
mercury or thallium on a plane.
Many possible mechanisms leading to stress corrosion cracking have been
found and they are either anodic or cathodic, i.e. the effect can be suppressed
in an electrolytic environment by having either the anode or the cathode on the
sample and the corroding agent as electrolytic medium. The most common cathodic
process is hydrogen embrittlement of metals in which at the crack tip the metal is
ionized due to charges that recombine with H + ions from the corroding medium
at the outer surface of the sample. The strength of the ionized metal is lower and
therefore it is more brittle. A typical anodic effect happens at metals that are
protected from the corroding agent by a coat of oxyde which is ripped open at the
crack tip exposing the pure metal to the agent just at the tipJ18^ In all cases the
velocity of the crack is controlled by the chemical reaction and is therefore rarely
greater than 0.1 m/s. This slow crack propagation can sometimes be observed on
the windshield of the car and also the crack shown in fig. 1.1c took several days to
be grown.

1.3.4 Statistical effects


For most materials the sample to sample variations of the brittle frac-
ture strength are very strong and a statistical treatment is useful. The most
commonly used form for the cumulative probability p of fracture is the "Weibull
distribution":^

ρ = 1-€χρ((σ-σ"Γ), (1.1)

where σ is the externally applied stress and ση is the largest stress at which the
probability of failure is zero and which is usually equal to or less than one-half the
value of the mean critical stress, ra, σ0 and au are material constants but they vary
substantially with the surface condition, sample preparation and temperature, up
to the point that for a given alumina m can take values between 3 and 12. In
fig. 1.9 we see experimental data for the distribution of strength for high-density
alumina specimens.
3. Phenomenology of fracture 11

I.U — i Π Γ^ i—
>
-
/S%

/z

0.8 ~"

0.6 -

- f% -

0.1. f% -.
•J
• /
• / A
• /
0.2

- Ä / A
n 1 J L_ 1

HO 160 160 200 220 ( M R a )


Strength

Figure 1.9 Distribution of strength values for the alumina Lucalox (dots) fitted by a Weibull
distribution of the form of eq. 1.1 (full line).
12 ff.J. Herrmann / Introduction to basic notions and facts

Figure 1.10 Young's experiment for (a) a homogeneous medium, (b) a medium with an ellipsoidal
pore and (c) a medium with a penny-shaped, thin microcrack.

1.4 The process of fracture


It is rather evident that fracture occurs through the propagation of cracks
but why do these cracks appear and how do they organize to break the system?
In the following we will describe qualitatively some of the most common processes
that yield to brittle fracture.

1.4.1 The behaviour of one microcavity


As already mentioned before real materials have plenty of heterogeneities
which can range from very localized interstitial vacancies to rather extended grain
boundaries. At length scales at which the medium can be approximated by an
elastic continuum we can disregard the detailed nature of these heterogeneities
and generically consider that there are randomly distributed "microcavities" in
the medium. Usually one distinguishes two types of microcavities: "pores" which
have roughly the same length in all directions and can be approximated by spheres
or ellipsoids, and "microcracks" which are typically 102 to 104 times longer than
wide, one says, they have an "aspect ratio of 102 to 10 4 ". In rocks, pores can be due
to intergranular spaces in sediments or to fluid inclusions in igneous rocks while
microcracks can be cleavages through mineral grains or lie inside grain boundaries.

The next useful step is to understand how a microcavity modifies the me-
chanical behaviour of a solid. So let us perform Young's experiment for a sample
with one cavity (figs. 1.10b and c) and compare it to a sample with no cavity
(fig. 1.10a). In the homogeneous sample of fig. 1.10a every volume element con-
tributes equally to the elastic response or more precisely, on any cubic volume
4. The process of fracture 13

element AV act three pairs of forces in opposite directions having for each vol-
ume element the same absolute value F. If AS is the surface of one face of AV
then σ = F/AS is the stress so that one can say that in fig. 1.10a the stress is
everywhere the same.
If one drills now a cavity into the medium those volume elements that would
have been inside the cavity cannot contribute anymore to the elastic response and
the remaining volume elements must take over the excess load, i.e. the stress that
would have acted on the missing volume elements. It happens, however, and this
is crucial for fracture, that the excess load is not shared equally by the remaining
volume elements but that the redistribution of stress is strongly inhomogeneous.
It is already intuitively clear that since at the height of the cavity (h in
fig. 1.10) the cross section is reduced the stress acting on volume elements at this
height h will be enhanced. But the increase of stress is not the same for all volume
elements on height h. If σο is the stress acting on the volume elements before
the cavity was drilled then the enhanced stress acting on volume element k is
Gk = Κτσ0 and we call KT the "stress concentration factor".
For a cylindrical hole in an infinite, linearly elastic medium it can be shown
analytically that next to the hole at the height h (point A in fig. 1.10b) one has
otk = 3 independent of the cross section while a* —> 1 for volume elements at
height h that are far from the hole. For an ellipsoid with the longer axis of length
2a perpendicular to the load one has

<** = 1 + 2 J ^ (1.2)

at point A in fig. 1.10b where p is the radius of curvature at point A.


For two-dimensional plates analytical calculations can be pushed much fur-
ther using conformal transformations in a complex formalism. So it is known that
for an elliptical hole in an infinite, linearly elastic medium at height h the excess
stress, Gk — σ 0 , decays like l/y/r close to point A while far from the hole the de-
cay is that of a dipole field, namely 1/r2 where r is the distance from point A.
More about this will be said in chapter 6. Close to A one can even calculate for
each of the three modes of fig. 1.7 the full dependence on the angle Θ as shown in
fig. 1.10b for each of the components of the stress. These formulas can be found in
the textbooks.[15J
Close to the tip of a microcrack like the one shown in fig. 1.10c the excess
load also decays in a plate like l/\/r. The proportionality constant AT, defined
through au — cr0 = K/y/2nrf(6) is called the "stress intensity factor" and it de-
pends on the length of the crack, the geometry of the sample, etc. The value of
K at which the crack breaks at the tip is the material-dependent constant, called
"toughness" Kc which was mentioned already before. For a given crack tip one has
three toughness values, one for each of the three modes of fig. 1.7. Knowing the
toughness of a sample is of more use to engineers than knowing just the breaking
14 HJ. Herrmann / Introduction to basic notions and facts

force in a specific experimental setup because the breaking force depends for in-
stance on the length of the crack, the geometry, etc. The typical toughness of steel
in mode I is ΚΪ€ = 5-10 GPa y/m.
We see from the enhancement factor of an elliptical hole at point A in
eq. 1.2 that the sharper the tip of a microcrack the more its stress is enhanced at
the tip. At an infinitely sharp tip the stress has in fact a singularity. So, even if the
externally applied stress σ0 is much below the breaking stress ac, locally, namely
close to the tips of microcracks, the stress can be much higher than ac and so the
material can break at these tips. Once a microcrack tears open at its tip it becomes
longer and therefore the stress at its tip is enhanced even more so that it will not
stop growing anymore. We see that within this picture crack tips are unstable.
In reality a mathematical divergence of the local stress at a cusp-like tip
does not exist. On one hand the continuum description is only valid up to a lower
cutoff at which the atomistic nature becomes visible. Consequently one cannot
get arbitrarily close to the crack tip neither can a tip become arbitrarily sharp.
On the other hand, if the value of the local stress passes beyond the yield point
it is certainly not allowed to consider anymore linear elasticity but in fact the
material becomes plastic. One therefore has the interesting scenario that although
the sample as a whole is elastic and breaks brittle there are regions inside the
material, namely around the tips of microcracks, in which the material deforms
plastically and breaks ductile. In fractography, the experimental technique that
investigates the physionomy of the crack surface, one can nicely visualize the two
types of behaviour occurring in the same sample.
The shape and the mechanical response of the plastic regions around crack
tips surrounded by an elastic medium is difficult to estimate. However, only the
characteristic size £ of these regions is essential for fracture because at £, which
is the shortest distance from the tip of the crack to the boundary of the plastic
zone, the singularity of the elastic stress field is cut off. The largest value ac of the
resulting stress field which is necessarily on this boundary has to be compared to
the toughness of the material if one wants to know when the crack will grow. In
reality there are even more effects that also introduce a cut off on the singularity of
the stress field, like viscoelasticity or disorder and these effects must also be taken
into account if one wants to know the critical stress. In numerical calculations such
a cut off is naturally introduced through the discretization of space. We will discuss
this point in more detail in chapter 5.
1.4.2 Interacting cracks
Up to now we have only discussed the behaviour of one single crack. In fact,
looking at fracture in the real world one gets the impression that most samples are
broken by just one crack. On the other hand, except when for test purposes a crack
is artificially nucleated by a notch before applying the load, in most materials no
particular crack can be singled out to be the critical crack that will break the
system and on which the toughness etc. could be measured locally. In fact we
4. The process of fracture 15

1
11
A

s * J
• ^ B
.

IF
Figure 1.11 Sample with an ensemble of microcavities of arbitrary shape randomly distributed
in space. At A two cracks attract each other, at B two cracks screen each other.

started out with a material full of pores or microcracks and therefore the next,
and perhaps essential point to understand fracture is to follow how out of many
essentially equivalent randomly distributed microcavities one crack will be born
that will tear the system apart. This enterprise is difficult and has been neglected
very much in the past but it is, at least for brittle fracture the way to understand
how the crack responsible for failure comes into existence.
Imagine an ensemble of microcavities of arbitrary shape and orientation
scattered randomly in space as shown in fig. 1.11. As discussed above each cavity
produces a stress enhancement field around it which depends on its geometry and
size and decays at large distances like a dipole field. Because of this slow decay
(slower than exponential) technically one has a long range interaction.
As long as one only considers linear elasticity the stress enhancement fields
coming from the various microcavities can be added together (superposition prin-
ciple) and one is left with an extremely complex stress enhancement field for the
ensemble which even numerically cannot always be handled. This problem will
come up again in several chapters (6, 7, 9) of this book.
The interplay of several stress enhancement fields can yield to the most
diverse effects: Between two close cracks that are aligned and perpendicular to the
load like the pair shown in fig. 1.11 at A the stress field is particularly enhanced.
For long, straight cracks one can calculate that at the tip of one single crack of
length a the enhancement is proportional to y/a while between two cracks of length
a the enhancement is proportional to a. Consequently the two cracks shown at A
in fig. 1.11 will attract each other and if the external load is large enough coalesce,
i.e.the neck between the two will break and they will become one single crack.
On the contrary, in other regions the stress field is substantially weakened
16 H.J. Herrmann / Introduction to basic notions and facts

like in the situation shown in fig. 1.11 at B. There two cracks are parallel to each
other and if the cracks are sufficiently long the stress between the two cracks will
be screened, i.e. the external load will not be felt. Prom screening to attraction all
scenarios are possible specially if more than two cracks are present.
The interesting question is, however, not just the stress distribution gen-
erated by the ensemble of microcracks but the growth and coalescence of these
microcracks when the external load is increased, i.e.its dynamical evolution. Sup-
pose that we increase the load from zero and reach the point when the first, i.e.the
most stressed region fails, for instance that two close crack tips coalesce. Due to
this coalescence the configuration of microcracks changes which for us means that
the stress enhancement field is modified. Although we only consider linear elastic-
ity the response of the system to the external load is now non-linear due to the
changes in stress enhancement each time locally a region fails.
Non-linear problems of this kind are of course very difficult to handle and
practically all that can be done are numerical simulations. Fortunately similar
problems of non-linearities due to growth in essentially linear media are known
from other areas, like viscous fingering in hydrodynamics or dielectric breakdown,
dendritic growth or electrodeposition (see chapter 8). Recently physicists have been
very interested in these problems ("growth models", "fractals") and a theoretical
framework (scaling laws, etc.) has evolved. The description of this framework is
one of the main tasks of this book.
While a precise calculation of the dynamical evolution of the ensemble of
microcracks is technically difficult it is easy to give a qualitative description: The
first regions that will fail, i.e. the most stressed ones, are determined by the initial
position of the microcracks, i.e. their random location. But once in a region a crack
has become larger and therefore the stress has been locally enhanced this region is
more susceptible for further cracking. So regions that failed once tend to fail again.
The effect of repeated failure will start to compete with the failure of the randomly
placed, stressed regions that did not yet fail. When the repeated failure takes over
one is roughly at the critical breaking load. The regions of multiple failure, i.e.
the growing cracks also compete against each other because the larger cracks grow
faster. Coalescence of cracks is the fastest mean of propagation. Finally one crack
wins and rushes through the system coalescing with the cracks it encounters on
its way. The time it takes for one crack to impose itself over the others strongly
varies with the configurations of microcracks and this explains the large sample to
sample variations encountered in the postfailure regime. The whole process can be
visualized quite nicely if one tears apart a piece of swiss cheese!
This picture which in this simplicity is only valid for brittle fracture does
of course not yet describe perfectly the reality because there exists a large variety
of additional effects inside the materials. Let us just mention two, namely "crack
arrest" and "healing". The "cohesive forces" which hold the material together and
which the growing crack must overcome have a certain range and tend to pull the
atoms together even if they are separated by a crack over several lattice units.
4. The process of fracture 17

Consequently if a crack does not open up very much its stress enhancement can be
substantially decreased even to the point of stopping the growth of a propagating
crack. The most common reason for crack arrest is, however, when the crack tip
encounters a hetereogeneity in the system. If the two sides of a crack are pressed
together the chemical bonds between the atoms can also be restored and a crack
can heal. Chemical reactions or crystallization inside the crack opening are even
more efficient mechanisms for healing. Both effects, crack arrest and healing, tend
to stabilize the process of fracture.

1.4.3 The energy of fracture


The description we gave of brittle fracture is rather detailed but not easy
to use for quantitative predictions. Alternative, more phenomenological approaches
to fracture have been developed and used over the years. The most popular one
concentrates on the energies involved in fracture and was first formulated in 1921
by Griffith J19^ The idea is based on energy conservation. When a load is applied on
a sample potential elastic energy is stored in the system. If a new crack is formed
or an existing crack grows part of this elastic energy is released. On the other hand
the formation or growth of a crack implies the creation of free surfaces, namely the
crack surfaces which are energetically less favorable than the bulk because chemical
bonds have to be broken up. So Griffith's arguments states that a crack grows if
and only if the release of potential energy is equal or larger than the surface energy
that is required for the crack to grow.
Neither the elastic nor the surface energy are easily accessible in experi-
ments or easy to calculate in general. But there are some cases in which calculations
are feasible like the case of one single elliptical flaw of length a perpendicular to
a uniaxial external displacement that is imposed on a plate (see fig. 1.10b). For
mode I rupture one finds in the limit of large p l a t e s ^ that the release of elastic en-
ergy dU caused by an elongation of the flaw by da is given by G = dU/da = πασ2/Ε
where σ is the externally applied stress and E is Young's modulus. G is called the
"energy release rate". According to Griffith's criterion the crack will grow if G
equals the surface energy needed to create two surfaces of length do because a
crack in a plate has two sides. In other words, the crack grows when G exceeds the
critical energy release rate Gc which is given by Gc = 2js where 7S is just the spe-
cific surface energy. Knowing the surface energy which is a material constant one
knows therefore the critical stress for mode I cracking. The critical energy release
rate Gc consequently plays a similar role as the toughness, i.e. depending on the
rupture mode and on the material it determines when a single crack grows. It is in
fact possible to write down a relation between the toughness and Gc (see chapter
5).
Up to this point it seems that we have just described in terms of energies
what we did before in terms of forces. Energies are, however, global quantities
while forces are locally measurable and instead of a local equilibrium of forces we
used the principle of energy conservation. Conceptually one is therefore tempted to
18 H.J. Herrmann / Introduction to basic notions and facts

generalize the Griffith criterion to arbitrary systems with many cracks and to use Gc
as a global, phenomenological parameter. The big drawback of this approach is that
due to various effects happening in real materials energy conservation is usually not
valid unless other contributions are taken into account. We already encountered the
example of viscoelasticity which comes from energy dissipation within the material.
Such a dissipation can be seen as an internal friction which can for instance heat
up the system. In the energy balance dissipation makes up for an additional, a
priori unknown term that must be treated as another phenomenological constant
and must be inserted by hand.
A particularly common loss of energy comes from the plastic regions around
the crack tips since every plastic deformation requires work. Again a phenomeno-
logical term, the "plastic deformation energy" 7 p can be introduced^21! and the
Griffith criterion for fracture is then modified to Gc = 7S + 7 P . Unfortunately in
many cases this is not just a small correction to the usual Griffith criterion because
for most metals and polymers 7 P is several orders of magnitude larger than %. It
seems nearly impossible to calculate 7 P or 7S from first principles.
Another interesting problem related to the non-local nature of energy con-
servation is energy transport. The elastic energy freed by the growth of the crack
is stored within a certain region and must be transported to the place where it
is needed. There are several possible mechanisms that can assure the transport of
elastic energy one being "phonons", i.e. waves inside the bulk like shear waves, but
the most common being "Rayleigh waves" ^ on the crack surface. Each mechanism
occurs with a certain velocity which for phonons is for instance the sound velocity,
i.e. (1.5 — 12) x 10 3 m/s and which for Rayleigh waves is about (1.5 — 4) x 10 3 m/s.
The existence of a finite speed for energy propagation introduces a time delay into
the energy balance which is important if very fast processes are considered. Into
the Griffith criterion these effects introduce an additional kinetic energy term.
K G > G c , i.e. if more energy is available than the minimum needed for
crack growth, the excess energy is often transformed into kinetic energy for the
propagating crack. Consequently the faster the crack grows the more excess energy
is available. We encountered this situation already when we discussed shocks. The
existence of a finite velocity v for energy transport imposes, however, a certain
limitation on this picture.
If the crack propagates faster than v the elastic energy which is released
behind the crack tip and which can therefore never reach the tip accumulates on
the sides of the crack. This accumulated energy tends to create more crack surface
leading to side branching of the crack. These side branches are usually short and
thus called "microbranching". If the velocity of the crack tip exceeds \/2 v the crack
tip can spontaneously bifurcate^ and two cracks simultaneously propagating with
speed v can be sustained as long as both are equally fast and one does not outrun
the other ("macrobranching"). If enough energy is available each tip can again
bifurcate and so one can finally have many simultaneously propagating tips. While
microbranching usually follows grain boundaries or cleavage planes macrobranches
5. Materials 19

can be intergranular.

1.5 Materials
The effects that we have discussed up to now, like plasticity, corrosion,
healing, etc., can in principle occur in all materials but their relative importance
can vary over many orders of magnitude. The origin for these large variations
can be found in the very different microscopic mechanisms that are involved in
the mechanical response. The resulting difference in the behaviour of the materials
makes up for the whole richness of fracture. In the following we will describe briefly
the characteristics of various materials.
Roughly speaking one can distinguish three families of materials studied
in fact by rather different communities: polymers, metals and what one may call
"rock-alikes". The last category would include materials like concrete, glass, ceram-
ics, heavy clay and tectonic plates. We will next give for each class some selected,
certainly very incomplete information. In the next chapter experts will give more
an insider view on some of the materials. The references will lead the reader to
more details.
1.5.1 Glass
The fracture of glass has fascinated researchers for a long time. It was
for glass that stress corrosion cracking^16! was first diagnosticized and that the
first theories for crack branching^23] were conceived. Typically glass has a strength
(breaking stress) under tension of 10-100 MPa but it is possible to prepare glasses
which are more resistent to tensile fracture than steel (10 GPa). It is interesting
to note that strength measured on supposedly identical samples shows already
a typical scatter of 10 to 20%. The slower the cracking process takes place the
stronger is the temperature dependence on the breaking strength as shown in
fig. 1.12. Until about 400 K the strength decreases with temperature due to "static
fatigue", i.e. in this case due to corrosion induced by the slow hydrolysis of the
silicates. At higher temperatures this process is counterbalanced by the desorption
of water and the strength increases again. Finally the temperatures are so high that
the glass fails by viscous flow. In the absence of an external stress the corrosion
can increase the strength^ as much as 60% because it blunts the cracks, i.e. it
increases the radius of curvature at the crack tip.
Fracture of glass nearly always starts at the surface and under tensile stress.
Therefore one can make glass more resistent to fracture by polishing its surface.
Abrasion, i.e. surface wear, on the contrary, makes glass more fragile. The crack usu-
ally nucleates at some manmade imperfection. Its subsequent propagation leaves
behind its fingerprint on the morphology of the crack surface. Around the initia-
tion point one sees a flat region, the "mirror zone", normal to the direction of the
tensile stress as shown in fig. 1.13a. The crossing of the rupture front with elastic
waves can leave behind ripples in the mirror zone called "Wallner lines". The mir-
ror zone is delimited by the "mist zone" which consists of fine stirations looking
20 H.J. Herrmann / Introduction to basic notions and facts

g I . i . i i i i i I
0 200 (.00 600 Θ00 K

Figure 1.12 Breaking strength σο of massive glass normalized by its value at 77 K as a function
of temperature for two different failure times.

like microscopic blades oblique to the crack plane. The mist zone appears when
the crack reaches speeds of about half the velocity of transverse elastic waves. It is
the onset to a region of crack bifurcation, surface roughening and even shattering
of the glass, called the "hackle zone", which occurs within a wedge-shaped chip
(fig. 1.13b). Shape and size of the zone can be related to the stress distribution^
and the depth of the mirror zone seems to go roughly like the inverse squared
tensile breaking strength. The speed of crack propagation can vary by about eight
orders of magnitude.

1.5.2 Rocks
Geology is particularly interested in the fracture of rocksJ25^ Evidently the
mechanical behaviour of rocks and minerals can be very varied. Let us just list some
typical values for the stress intensity resistance in mode I measured in MPa y/rn:
0.01 to 0.3 for coal, 0.2 for calcite, 2.4 for natural quartz and about 3 for granite and
sandstone. In fact, these values are strongly dependent on crack velocity, moisture,
temperature, pressure and the geometry of the performed test as illustrated in the
example of fig. 1.14. In addition, rocks can be very anisotropic. For example, the
tensile strength of shale measured normal or perpendicular to the bedding may
vary by a factor ten, the compressive strengths vary less.
The most common tests on rocks are performed under compression because
on one hand this mimicks the enormous pressures to which rocks are submitted
deep inside the earth's crust and on the other hand because of important techniques
like hydraulic fracturing or deep explosions. Hydraulic fracturing in which fluid is
injected into the rock under pressure is a way to improve the flow of oil and
explosions are used in seismic measurements. The brittle to ductile transition was
also discovered^ in compression experiments.
5. Materials 21

Hackle

Mist

Initial flaw

(a) (b)

Figure 1.13 Morphology of a crack through glass; (a) top-view of the crack surface, (b) three-
dimensional view of the wedge-shaped chip of hackle.

Compression experiments are often performed by applying a "confining


pressure" σχ from all sides and putting in addition a "differential stress" σ 3 — σχ
in the z-directiW 11 } (see also fig. 1.4b where <r3 — σχ is called F/A). For σχ = 0
igneous rocks typically break at σ 3 = 100-200 MPa while high porosity sediments
break at about σ 3 = 10 MPa. If the confining pressure is increased the material
becomes tougher. This can be seen for igneous rocks in fig. 1.15a which shows the
increase of the critical differential stress as a function of σχ.
The applied differential stress can also be a shear and in this case the
confining pressure increases the strength of the sample even more than in the
case of compression. Under shear the cracks have an inclination with respect to
the z-axis and the angle φ of this inclination depends on the angle Θ between
the force F of fig. 1.4b and the 2-axis as illustrated in fig. 1.15b. Shear fracture
under a confining pressure is encountered when a tectonic fault breaks up and
starts moving which is the principal source for earthquakes. The sliding of the
two faces of the crack surface on each other in the presence of pressure is called
"slip". The asperities of the crack surfaces introduce a friction force during the
slip motion. This friction decreases with the amount of sliding AL/L, an effect
called "slip-weakening", because the asperities are sheared off and crushed by the
motion. The resulting debris found inside the crack is called "gouge" in the case
of tectonic faults. Moving like an avalanche of sand ("cataclastic flow") gouge can
act as a lubrifying fluid between the faces of the crack allowing for the creep of
faults. Theoretical understanding of this process is still very poor. In fig. 1.16 we
see slip-weakening measured on granite for various confining pressures.
22 H.J. Herrmann / Introduction to basic notions and facts

i
u •

water ♦ waterί
J
>air / // 'J" ·''
/
5 r 20°C ♦ 60°C i /30°*>RH ' / / .*

r A
/ /20°C
♦ / /
;
/ '
/zo-c j %·
£ 6 ♦ 80°c/
1 t
V v
>N
/ · /20°C
o
; 7
. ;
/
♦ ♦
; Arkansas /
novaculite / // ' 7i
water

σ
t_
ϊ' / /
/ / /

/
I

" 8 / /
o
C7>
Tennessee sands fone Mojave
O
quartzite
1 / /
9 \ //
water

1 1 1 1 1 1 1 1 11 1 1 1
0.1 0.2 0Λ 0.6 0.8 1
Kj (MPam 1/ 2)

Figure 1.14 Log-log plot of the stress intensity factor K\ in mode I as a function of the crack
speed for different quartz rocks at different temperatures measured either in water or in air (taken
from ref. [25]).

60° ~ 90°
Θ
(b)
Figure 1.15 Typical behaviour of igneous rocks, (a) Dependence of the differential fracture stress
03 — &i at which the system breaks on the confining pressure σ\ in compression experiments, (b)
Dependence of the inclination angle φ of the crack on the angle Θ that the applied force forms
with the 2-axis under shear.
5. Materials 23

σ
3'σ1

1.0

0.8

0.6

0.4

0.2

0 0.01 0.02 0.03 AL '


Figure 1.16 Slip-weakening measured on initially intact granite for various confining pressures
σ\. We show the differential shear load σ$ — σ\ as a function of the amount of sliding AL/L. The
slip motion sets is once the maximum of the curve is reached.

Once the moving force behind the slip motion is gone because the stress
that had built up in the tectonic plates has been relaxed by the motion of the
fault, the crack surfaces stop moving, i.e. they "stick". Due to the drift of tectonic
plates the stress slowly builds up again and the resulting "stick-slip" motion can
be very complex. Already one-dimensional, deterministic models for stick-slip can
show chaotic behaviour^ due to the decrease of the frictional force with the slid-
ing velocity. A consequence of these simple models is not only the unpredictability
of earthquakes but also a power-law distribution for the lengths and the amount
of sliding AL/L of individual slipsJ26^ Stick-slip motion is also encountered when
two pieces of rubber are moved against each other under pressure. The stick-slip
behaviour can be suppressed either by lowering the confining pressure or by increas-
ing the temperature. If a fault sticks long enough it can heal under the enormous
pressures that exists in the earth's crust due to a process that ressembles the for-
mation of sedimentary rocks. The recracking of healed cracks which is common in
fault zones is a difficult theoretical problem.
Laboratory tests on rocks are not simple. Often rather big samples are
needed in order to avoid too large sample to sample variations because the het-
erogeneities in rocks can reach the size of centimeters. Another typical problem
encountered by geologists is that when a sample is removed from its in situ stress
environment fractures can develop inside changing its behaviour in a test. Uniaxial
24 H.J. Herrmann / Introduction to basic notions and facts

tensile strength is not easy to measure because in this case it is difficult to maintain
the external load uniform.
Microscopically the typical defects that nucleate cracks in rocks have an
opening but no shear displacement and only propagate in mode I. They can be
"joints" that are formed during the cooling of igneous rocks, the drying of sediments
or as tension gashes during faulting. They can also be about 1000 μτη long and 1 μτη
wide penny-shaped microcracks within grain boundaries or intergranular cleavage
planes.
These microscopic flaws propagate due to mechanical deformations but also
due to thermal expansion, both common in the earth's crust. While propagating in
mode I the surface of the cracks shows characteristic tilt and twist configurations
due to the existence of shear fields. On larger scales mode I fracture in rocks
also shows the mirror-mist-hackle scenario described above for glass. If a crack
has grown long enough it can also propagate in mode II or mode III. In this
case the surface morphology is less unique: In some rocks like sandstones one
sees deformation bands, in other cases the crack is accompanied by plumes of
microcracks called "feather fractures".
There is a delicate balance between the stress required to cause a mineral
grain to cleave and the stress required to cause brittle grain boundary cracking.
This balance can be upset by small changes in temperature, impurity content and
rock texture. The fracture path can therefore be quite complex on large scales.
The presence of fluids, usually water, can have various effects. The high
pressure of fluids enclosed in the pores of igneous rocks can produce strong local
stress fields and the presence of such local constraints is k n o w n ^ to change the
elastic response of a system. They evidently also have an important influence on
the brittle fracture behaviour. Healing is another effect that typically requires the
presence of fluids. Hydrothermal processes including precipitation and solution
under pressure can rapidly fill open microcracks with "cement" J28^
Fracture of rocks is usually brittle. But we already saw that under pressure
a transition to ductility can occur.^ Geologists concerned with deformation and
failure of the earth's crust, consider pressures as high as 5 GPa when sedimentary
rocks are plastic while igneous rocks behave relatively elastic. High temperatures
also enhance the tendency to ductility. Two mechanisms can sustain plastic flow in
rocks. On one hand one can have slip and twinning of crystal planes on each other
giving rise to crystal plasticity. On the other hand one can have the microscopically
brittle mechanism of cataclastic flow where the material slides on the debris of local
fragmentation. Fragmentation will be the subject of the last chapter of this book.
1.5.3 Ceramics
Ceramics is a large class of man-made materials to which, in an extended
sense, also belong glass and concrete. What crack resistance is concerned the spec-
trum is very wide for ceramicsS29^ There exist some heat insulators that can be
crushed between the fingers while a mixture of titanium boride and boron carbide
5. Materials 25

1/7d
Figure 1.17 Typical behaviour of the breaking strength versus the reciprocal square root of the
average grain diameter d for alumina and magnesia (A) and for beryllia (B).

can attain a compressive strength of 4 GPa; porcelain has 250-350 MPa. Tensile
strengths are typically eight to ten times lower.
Usually ceramics are brittle and in this case, like in glass, cracks are gen-
erally initiated at surface flaws. Consequently the strength of brittle ceramics can
be increased reducing the presence of these flaws for instance by flame-polishing
or etching the surface. For a certain class of ceramics, called alumina (AI2O3), the
strength can be varied through surface treatment between 0.5 GPa and 15 GPa.
We saw in a previous section that an increase in the strength can lead from
brittle to ductile rupture. This is the origin of the "Joffe effect" of alkali halides:^
they are brittle in air and ductile in water. The water polishes away surface imper-
fections of the sample by solution increasing the strength and rendering it ductile.
When the sample is then dried small crystallites precipitate on its surface lowering
the strength again and it rebrittles. Materials that behave in such a way are often
called "semibrittle" because their yield strength and their fracture strength are
close. Most ceramics are hard and are semibrittle only at high temperatures. Soft
ceramics can be semibrittle at low temperatures. Ductility at low temperatures
occurs for silver chloride and bromide but is generally very rare for ceramics. The
dominant mechanism in ductile rupture is crystal plasticity.
A typical problem of material science is to increase the toughness of a
given ceramics. We already discussed the effects of surface treatment. Another
possibility is to change the internal structure of the material. In fig. 1.17 we see
schematically how the breaking strength depends on the average grain diameter d
for various oxides. Two different behaviours are observed, the transition between
the two being temperature dependent. At small grain sizes intergranular cracking
26 H. J. Herrmann / Introduction to basic notions and facts

800

600

in
Φ (,00
<7)
200

0 2 I 6 Ö 10 12 14
Elongation
Figure 1.18 Breaking characteristics for glass and carbon fibre reinforced glass (CRG) taken
from ref. [29].

dominates and for large grains one has mainly intragranular effects.
The probably most efficient way to increase the toughness is mixing var-
ious components, i.e. to make "composites". Concrete is a well-known example.
Spectacular are also the fracture strengths of more than 1 GPa for mixtures of
ceramics and metals, called "cermets". The reinforcing of ceramics with fibers is
another technique which not only increases the breaking strength but also changes
the post-failure behaviour. This is illustrated in fig. 1.18 for glass reinforced with
carbon fibers. The carbon fibers avoid catastrophic failure because they often do
not break where the glass breaks but have to be pulled out of the glass by the
opening crack. More details will be given about these problems in the sections on
ceramics, composites and concrete of the next chapter and models for the rupture
of carbon fibers will appear in chapter 6.

1.5.4 Polymers
In contrast to rock-alikes, polymers and metals are typically ductile and
for this reason beyond the scope of most statistical models of this book with the
exception of chapter 9. Their behaviour is, however, so rich that we will discuss
some aspects here and two sections will be dedicated to them in the next chapter.
The rupture of polymers^32! is usually and most easily studied under tension
as the reader can experiment with a rubber band. Their failure behaviour can
depend very much on temperature as shown schematically in fig. 1.19. At very
low temperatures they fail brittle (a) and when the temperature is increased a
transition to ductility takes place (b) with the usual necking instability occurring
when the "ultimate stress" is reached. A neck can be initiated either because at
some point the cross section of the sample is a little smaller or because a fluctuation
in the material properties may cause a local reduction of the yield stress.
5. Materials 27

Figure 1.19 Breaking characteristics for a typical polymer tested at different temperatures: (a)
brittle, (b) ductile, (c) necking and cold-drawing and (d) rubber-like behaviour.

The higher the temperature the more the material can be stretched and the
more the polymer chains become oriented in particular in the neck region. This
orientation produces strain-hardening which stiffens the material and stabilizes
the neck again. As a consequence the neck propagates along the length of the
sample instead of strangling it, an effect called "cold-drawing" (c). In this situation
elongations of 300-1000% can be achieved before rupture.
Finally at even higher temperatures the polymer undergoes a transition
(glass transition) to rubber-like behaviour (d) in which the material can be elas-
tically stretched over several times its size. Here rupture can be understood by
visualizing the rubber as a network of crosslinked polymer chains similar to a ran-
dom fabric. Small cracks nucleate at the weakest crosslinks until one dominating
crack tears the fabric apart.
Since polymers are usually viscoplastic their rupture behaviour strongly
depends on the "strain rate" έ, i.e. the speed with which the elongation takes
place. In fig. 1.20 we see how the tensile strength ac and the "ultimate strain"
e u , i.e. the elongation at which necking sets in, depend on έ for a certain type of
rubber.
An interesting phenomenon that often precurses polymer fracture is the
appearance of "craze" J33^ Craze is a dense network of fine planar defects often
starting at the surface of the sample. These defects lie orthogonal to the direction
of the tension. They contain about 50% free space and the rest are oriented polymer
chains cutting through the planes of these defects. The formation of craze can be
28 HJ. Herrmann / Introduction to basic notions and facts

100 ί-
OJ 50
a. 13
Σ
fl>

10
υ
D 5 \
JO.
■4·*
O) c
c
Φ

<
Φ
>
/ 1
m
c
.5
.<L> =zZ- I I I I I I 1 I

-8 -4 0 4 8
-log E a T

Figure 1.20 Variation of the tensile strength ac (full line) and the ultimate strain eu (dotted line)
of a rubber with strain rate i. The strain rate is in fact multiplied with a temperature-dependent
factor ατ such that the data points fall on the shown curves for all temperatures T (see ref. [32]).

explained by what is called "cavitation", namely a spontaneous appearance of voids


that happens in some plastic media under tensile stress.
Under cyclic fatigue^ heat is generated in polymers due to the relative
motion of adjacent chains. Since polymers are bad heat conductors the temperature
of the sample rises. The value T of this temperature depends on the frequency
as shown in fig. 1.21 for polyethylene. The temperature rise can lead to thermal
softening failures.

1.5.5 Metals
Because of the technological importance of metals their fracture behaviour
has been studied most intensely^ also at rather microscopic levels. The clue to
the understanding of metal failure is the motion of dislocations.
"Slip", i.e. the translation by a multiple of the lattice constant of one crystal
plane on the other is equivalent to the motion of an edge dislocation along the
plane. Since real metal grains are usually full of interpenetrating dislocations many
complex relative motions are possible inside a grain, giving rise to the mechanism
that underlies plasticity. Dislocations can be locked by present impurities in the
form of a slip band, called "Lüders band". If the stress is increased the dislocations
or a Lüders band can unlock and sweep through the grain generating plastic flow.
Under sufficiently high stress also "twin", i.e. relative motions of crystal planes by
fractions of units of the lattice constant can occur.
5. Materials 29

10 20 30 t0 t[minj

Figure 1.21 Specimen temperature T as a function of time t during cyclic fatigue of branched
polyethylene for various frequencies ω.

The collective motion of dislocations produced by the deformation of the


metal grains can generate interesting spatial patterns like regular networks of dis-
location bands or a self-similar decomposition of the Lüders bands into fine sub-
bands. A continuum description of the motion of the density of dislocations in
reaction-diffusion equations can explain some of these patternsJ36^
Due to the mechanisms that we described one of the keys to increase the
strength of metals is controlling the impurities. Strong steel is obtained from iron
by adding carbon and certain metals. The tensile strength of a piano wire can reach
300 GPa which is, however, still one order of magnitude smaller than that of the
strongest silica fibers. Impurities can also substantially lower the strength of metals
as happens in the hydrogen embrittlement mentioned in section 3 or in the "blue
brittleness" produced by the presence of nitrogen in carbon steel. Consequently
the strength of alloys of iron and nitrogen can be as low as 0.5 GPa.
The list of materials for which fracture has been studied is by far not ex-
hausted. The fibrous and cellular structure of biological substances^13! like w o o d ^
and bone^ can have rather exceptional breaking properties. The tearing of paper
and textiles is of interest to certain industries and the rupture of ice^ has been
investigated by geoscientists.
30 H.J. Herrmann / Introduction to basic notions and facts

1.6 Outlook
We have given a very brief and therefore superficial overview of the tra-
ditional knowledge about fracture. The aim of this chapter was to give the non-
specialist reader a feeling for the richness of the field and to guide him through the
current terminology.
With the exception of the next two chapters the rest of this book is not
intended to work out the details of what was said here but to present less traditional
approaches. In fact, most of the material presented in this chapter is rather well-
established as can be seen from the age of many of the references. Nearly all the
subjects that have been touched here are elaborated pedagogically and in detail in
many textbooks, some of them having been listed as references.

I thank D. Frangois and D. Stauffer for critical reading of the manuscript.

References
1. R.W. Hertzberg, Deformation and Fracture Mechanics of Engineering Ma-
terials, (John Wiley, New York, 1976) or Aval de S. Jayatilaka Fracture of
Engineering Brittle Material, (Applied Sei. Publ., London, 1979); other good
textbooks on fracture are refs. 11 and 15.
2. H. Liebowitz (ed.), Fracture, Vols.I-VII, (Academic Press, New York, 1986)
3. Stress Corrosion Cracking of Metals: A State of the Art, Symp. of the Am.
Soc. for Metals, (Detroit, 1971)
4. R. Englman and Z. Jaeger (eds.), Fragmentation, Form and Flow in Fractured
Media, Ann. Israel Phys. Soc, Vol.8, (Neve Ilan, 1986)
5. S. Timoshenko and J.N. Goodier, Theory of Elasticity, (McGraw-Hill, New
York, 1951)
6. T. Travers, D. Bideau, J.P. Troadec and J.C. Messager, J. Phys.A 19, L1033
(1986) and H.J. Herrmann, D. Stauffer and S. Roux, Europhys. Lett. 3, 265
(1987)
7. W.R. Schowalter, Mechanics of Non-Newtonian ßuids, (Pergamon, 1979)
8. L.D. Landau and E.M. Lifshitz, Elasticity, (Pergamon, 1960), p.130
9. T. von Kärman, Z. Ver. Dt. Ing. 55, 1749 (1911)
10. J.A. Hudson, S.L. Crouch and C. Fairhurst, Eng. Geol. 6, 155 (1972)
11. M.S. Paterson, Experimental Rock Deformation: The Brittle Field, (Springer,
1972)
12. Z.T. Bieniawski, in Structures Solid Mechanics and Engineering Design Teeni
(ed.), (London, 1971)
13. L. J. Gibson and M.F. Ashby, Cellular Solids: Structure and Properties, (Perg-
amon, 1988)
14. see e.g. M. Larsson, A. Needleman, V. Tvergaard and B. Storakers, J. Mech.
Phys. Solids 30, 121 (1982)
References 31

15. D. Broek, Elementary Engineering Fracture Mechanics, (Martinus Nijhoff, Dor-


drecht, 1986)
16. T.C. Baker and F.W. Preston, J. Appl. Phys.17, 179 (1946) and M. Grenet,
Bull. Soc. d'Encouragement 4, 838 (1899)
17. see e.g. M.G. Fontana in Stress Corrosion Cracking in Aircraft Structural Ma-
terial, (NATO, AGARD Conf. Proc. Ser. No. 18, 1967)
18. H.L. Logan, J. Res. Nat. Bur. Stand.48, 99 (1952)
19. A.A. Griffith, Phil. Trans. Roy. Soc. of London 221, 163 (1921)
20. C.E. Inglis, Trans. Inst. Naval Architects 55, 219 (1913)
21. G.R. Irwin, Fracturing of Metals, ASM publ., p.147 (1948) and E. Orowan,
Welding Journal 34, 157s (1955)
22. Lord Rayleigh, Proc. Lond. Math. Soc. 17, 4 (1885) and R.M. White, Proc.
IEEE 58, 1238 (1970)
23. E.H. Yoffe, Phil. Mag. 42, 739 (1951)
24. J.W. Johnson and D.G. HoUoway, Phil. Mag. 14, 731 (1966) and E.H. Andrews,
J. Appl. Phys. 30, 740 (1959)
25. B.K. Atkinson, Fracture Mechanics of Rocks, (Acad. Press Geology Series,
1987)
26. R. Burrigde and L. Knopoff, Bull. Seismol. Soc. Amer. 57, 341 (1967) and J.M.
Carlson and J.S. Langer, Phys. Rev. Lett. 62, 2632 (1989)
27. S. Alexander, J. Physique Lett. 45, 1939 (1984)
28. E.R. Padovani, S.B. Shirey and G. Simmons, J. Geophys. Res. 87, 8605 (1982)
29. R.W. Davidge, Mechanical Behaviour of Ceramics, (Cambridge Sol. Stat. Sei.
Ser., Camb. Univ. Press, 1979)
30. A. Joffe, M.N. Kirpitschewa and M.A. Lewitsky, Z. Phys. 22, 286 (1924)
31. W. Weibull, Kg. Svenska Vetanskapsakad Handl. 151, (Stockholm, 1939)
32. I.M. Ward, Mechanical Properties of Solid Polymers, (John Wiley, 1971), p.367
and T.L. Smith, J. Polym. Sei. 32, 99 (1958)
33. S. Rabinowitz and P. Beardmore, CRC Crit. Rev. in Macromol. Sei. 1,1 (1972)
34. J.M. Schultz, in Treatise on Materials Science and Technology, Vol. 10 ed. J.M.
Schultz, p.599 (Acad. Press, 1977)
35. D. McLean, Mechanical Properties of Metals, (Krieger Publ., Huntington,
N.Y., 1977) and W.A. Wood, The Study of Metal Structures and their Me-
chanical Properties, (Pergamon, 1971)
36. E.C. Aifantis, J. Eng. Mat. Tech. 106, 326 (1986)
37. C. de Zeeuw, in Modern Materials Vol. 1, ed. H.H. Hausner (Academic Press,
1958)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
©Elsevier Science Publishers B.V. (North-Holland), 1990 33

2. Experimental evidences for various materials

2.1 Rupture and deformation of ceramics


Jean-Louis Chermant^

2.1.1 Introduction
Ceramics are among the most advanced materials presently available for
high technology applications (fig. 2.1.1).^ Although there is no clear-cut boundary
between ceramics and metals, a tentative definition has been set forward by the
British Ceramic Society^ as "all solid manufactured materials or products that
are chemically inorganic, except for metals and their alloys, and which are usu-
ally rendered serviceable through high temperature processing". These are gener-
ally borides, carbides, halides, nitrides, oxides, silicides... and cermets (ceramic-
metals). There are many kinds of ceramics with specific functions in a wide variety
of fields. "Fine ceramics" generally designates ceramics which possess high additive
value: that is the case for electrical, insulating, magnetic, optical, superconducting,
thermal and thermomechanical ceramics, by contrast to "traditional" ceramics in
use in the building industry or as classical refractories.
2.1.2 Structure and bonding
Ceramics display inorganic crystal structures or, occasionally, glass-like
amorphous structures. According to the chemical species involved, the bonds may
be ionic, covalent or metallic. Additional van der Waals bonds also occur.^ Pro-
cesses and properties will depend on the nature of the bonds and the structure.
Ionic bonding occurs when one atom gives up one or more electrons which
are accepted by one or several other atoms in such way that an overall electrical
neutrality is maintained. Each atom thus gets a stable, filled electron shell. Metallic
oxides and halides have mainly ionic bonding (ionic structure of rock-salt type).
Covalent bonding occurs when two or more atoms share electrons. Here
again, this generally endows each atom with a stable, filled electron shell. Un-
like metallic and ionic bonds, covalent bonds are directional, each covalent bond
t LERMAT, URA CNRS 1317, ISMRA, Campus Univ.II, Bd. Mal. Juin, F-14032 Caen Cedex,
France.
34 J.-L. Chermant / Experimental evidences for various materials

t
O
o

1300

1200

1100

1000

900
800
1960 1970 1980 1990 2000 2010

Figure 2.1.1 Change in the temperature in use for different classes of materials as a function
of years from 1960: SA: superalloys; DS: metals strengthened by directional solidification; CVD:
materials protected by a chemical vapor deposition coating; CMC: ceramic matrix composite
materials (according to Clark and Flemings^).

involving a pair of electrons shared between two nuclei, so that, at least in the
simplest cases, the probability distribution for each electron may be visualized by
an electron cloud with a relatively high density along the axis joining the nuclei
(or close to it). Carbides and nitrides (or silicon) have mainly covalent bonding.
This implies that, in contrast to ionic ceramics, their crystallographic structure is
not governed principally by atomic (or ionic) size criteria.
Metallic bonding is, of course, typical of metals, but it can also appear
in ceramics: for example, metal transition carbides display bonds with partially
metallic character. This means that valence electrons are freely shared by all the
atoms in the structure.
Lastly the van der Waals bonds are due to electrical dipole interaction,
these dipole being either permanent or induced. In view of the energy involved,
van der Waals interactions may be considered as of only secondary importance
with respect to their effects on the structure and properties of the materials which
are of concern to us here.

2.1.3 Ceramic processes


Although ceramics have been obtained for several thousand years by knead-
ing wet clay followed by sun-drying, they are nowadays manufactured mainly by
sintering processes, either in liquid or solid phase, or by chemical vapor deposi-
tionJ 3-5 ) Sintering, where particles are agglomerated below the melting tempera-
1. Rupture and deformation of ceramics 35

Figure 2.1.2 Transmission electron micrographs showing the presence (a) or absence (b) of
secondary phases in grain boundaries: a) SiC - 0.75A1; b) SiC - B.

ture, can be achieved with pressure ("H.P.", hot pressing or "H.I.P.", hot isostatic
pressing) or without (pressureless sintering). It may also rely on a chemical reac-
tion: this is the so-called reaction sintering, leading, for example, to reaction-bonded
silicon nitride (RBSN) or to reaction-sintered silicon carbide (RSSC). Sometimes
other very specific methods can also be used, such as dynamic compaction or elec-
tromelting.
Depending on the ceramic type, the shape and the required characteristics,
numerous forming processes can be resorted to. The main ones are: die pressing,
slip casting, injection molding and extrusion molding.
Due to the covalent character of the bonds for many of these ceramics, solid
state sintering is very difficult to achieve without high pressure. Beside classical
binders and lubricants used during any sintering process, some additives with ten-
sioactive properties are required for ceramic sintering. These additives are of two
types:
- those which lead systematically to the formation of a vitrous phase: such is the
case for AI2O3, CaO, Ln 2 03, MgO,... additives; one may even use aluminium
since at high temperature it is readily oxidized and leads then to a glassy phase
too (fig. 2.1.2a),
- those which do not lead to the formation of a glassy phase: such is the case for
boron additives (fig. 2.1.2b).

2.1.4 Ceramic morphology


Ceramic morphology will depend on the ceramic powder characteristics and
of the type of process used.
On the one hand, one can observe morphological evolutions related to the
characteristics of the grain boundaries, where the glassy phase is located (or at the
36 J.-L. Chermant / Experimental evidences for various materials

Figure 2.1.3 Optical micrographs of silicon carbide obtained by hot press sintering, with boron
additive (from CGE Company) (a) and with 0.75% Al (from Elektroschmelzwerk Company) (b).

triple points) and where impurity concentrations arise.


On the other hand, by changing the characteristics of the initial powder
and the process parameters, ^ a given type of ceramic can assume several types of
morphology characterized by the size of crystals, their distribution(s), their shape,
their mean contact number, and so on. These are the features that one usually has
in mind when one speaks of disordered materials or systems.
As we know that the physical properties can depend on the microstructure,
which in turn depends on the process involved, it is more important to control
the microstructure for these materials as closely as possible than it is for metals.
That can be achieved by tayloring the chemical and physical characteristics of the
ceramic powders and the process parameters. The final product can be controlled
by automatic image analysis.^ As an illustration, figure 2.1.3 displays micrographs
of two different batches of SiC.

2.1.5 Rupture behaviour


Compared with metals, ceramic materials show high elastic moduli, ranging
from 70 to 400 GPa, which decrease very slowly with increasing temperature, at
least in the range 20°C to 120°C.[3'5]
Due to the high strength of the bonds in ceramic materials, one consid-
ers their theoretical strength, a t h, which is defined as the tensile stress typically
required to break atomic bonds. It ranges from 1/10 to 1/5 of the value of the
elastic modulus, and the real rupture stress for polycrystalline ceramic lies within
the range of 1/100 to 1/10.
The high value of σ^ is related to the high energy of the atomic bonds in
these materials, resulting in a high value of the fracture surface energy, 7, in the
1. Rupture and deformation of ceramics 37

expression of ath (see also eq. 5.2.5):

0"th (2.1)
ÜQ

where E is the elastic modulus and α$ the interatomic spacing.


The theoretical strength is rarely obtained in engineering practice (except
in some cases with whiskers or monofilaments), because of the presence of defect
flaws (either structural or due to fabrication), which results in stress concentrations:
fracture then arises at a load well below the theoretical strength.
It is possible to give an interpretation of the change in the rupture stress
with temperature for different materials, keeping in mind that these values depend
not only of a t h, but also on the mean grain size, porosities, shape of crystals,
etc. (fig. 2.1.4). We observe two types of evolutions, which reflect microstructural
differences:
- a decrease starting in the region 600-800°C, mainly for nitrides (and also sialons,
i.e. ceramics made of Si, Al, 0 and N), due to the change in the viscosity of the
glassy phase at the grain boundaries,
- a constant value or a slight increase up to 1300-1400°C, followed by a decrease;
this behaviour is specific for materials without a secondary phase at the grain
boundaries.
The high sensitivity of these materials to defects, as can be seen from figure
2.1.5,^ is a clear illustration of the fact that the rupture stress is not an intrinsic
parameter. It will depend on the porosities (size and shape) as well as inclusions
and the pore-crack combination. But it is also determined by the critical size defect,
i.e. the critical size leading to the catastrophic failure at a given stress, which, for
ceramics, can reach the size of one crystal. So we can use the stress intensity factor,
ÜTi, or the "R-value", which takes into account the increasing size, a, of the crack
during the loading, when there is stable crack propagation. This approach considers
fracture in terms of crack surface displacements and stresses at the crack tip, and
not in terms of critical flaw size.
It has been demonstrated by Neuber^ and more recently by BureschJ10,11^
that the fracture of certain ceramics is determined by the critical values for the
notch fracture stress, a mc , and the size of a process zone, p c , (fig. 2.1.6). This process
zone is analogous to the plastic zone of metals, in the sense that the microcracked
region in the immediate vicinity of a crack tip is responsible for the non-linear
behaviour of ceramics. In the process zone, there exists a constant stress, a mc , which
may be interpreted as the stress necessary to break either the grain boundaries or
the crystals themselves. This stress depends on both the theoretical stress in a
hypothetical ceramic without defects and the residual stresses in the real material.
When the mean value of the stress in the process zone, a m , reaches the critical
value a mc , in a critical volume, instability occurs and the main crack propagates.
The value of the critical stress intensity factor in mode I (opening mode) is then
38 J.-L. Cheimant / Experimental evidences for various materials

t Si3N4HP13Y^2A, _ ^ SiC

E Si 3 N 4
z \ — _ IN 100
\ I
b
C^-^\ A Al 2 0 3
SiCHP + A t γ
Λ \ \ Υ2Ο3
600
c . r uDp \ \ X
400 <RB SiC (Rofol) \ U ^ 1
V-v
R
SiCaFN \
\ "· ^
400

300

200 ' · v. I
100

I 1 1 L·— 1 1 1 1 J
400 800 1200 T°C-

Figure 2.1.4 Change in the rupture stress in 3-point bending for different ceramics as a function
of temperature. For comparison, the change for a classical superalloy (IN 100) is also given.

J|
1 HARTLEY 1982
E
z
Σ
600 ^ s .

400

- Al203

200 , «j ■ ■ I
25 50 75 'μιτι*

Figure 2.1.5 Change in the rupture stress, σΓ, by 3-point bending of alumina containing defects
of different sizes, a, (according to Hartley^).
L Rupture and deformation of ceramics 39

^Q|

/ ° m c = /=·_

IC

/
1/ 1 -►
,"2
y Original crack tip position
r > r 0 =>p>p c => K Q > K I C
2 Wake of the microstructural
region
3 Microcracked region r s r 0 o p = P c * KQ =KIC

4 Process zone
*<r0 x> p = p c = > K Q =KIC
5 Elastic region

Figure 2.1.6 The different regions in a plate containing an advancing crack. Relationship between
the crack tip radius and the process zone radius.

given by
(2.2)
Pc> 0"mc and Kic are considered as characteristic parameters of the material.
In many cases, particularly when ceramics contain a secondary phase, a
slow crack growth arises under load in a domain where the stress intensity factor
is lower than ÜTicJ12'13^ Then it is possible to relate the stress intensity factor to
the crack velocity, v = da/dt
v = AK1}, (2.3)
A and n being constants.
2.1.6 Strength and statistics
If enough specimens are tested, it is generally found that the Weibull distri-
bution function describes correctly the results J14-16^ Moreover assuming the defect
distribution is uniform throughout the material, it is obvious that the chance is
higher to reach the critical conditions somewhere in the specimen when its vol-
ume is greater. This so-called volume effect has been dealt with quantitatively by
Weibull, who demonstrated that the values a(Vi) and σ ( ^ ) of the rupture stress
of two specimens with respective volumes V\ and V^ are related by the following
formula, where m is called the Weibull parameter:
1/m
(2.4)
σ(Υ2) WA
40 J.-L. Chermant / Experimental evidences for various materials

4 99
" GOVILA-1983 ti
1 95 if 7
^90 II
"-80 MLE /
70 2
60 _ cal.E lcm iZ

Γ
50
40
30 _ aw=750MNm-2 / I I aw
/ j
347MNm-2J \
=
if a F = 363MNm-2
20
im
#y787MNm

10 'Γ
7 ·
6 u #/ II
ih II ·
4 II II
II II
II
2 Si 3 N 4 -MgO
a 20°C b1371°C
1 1 1 1 1 1 1
0 ' l I I l l I I. I I I
150 200 400 600 800 1000 150 200 400 600 800
CTMNm-2_φ.

Figure 2.1.7 Results from the Weibull tests on Si3N4 - MgO (Norton, NC 132) in 3-point
bending, at 20°C and 1371°C. The full lines correspond to results based on the linear elastic
fracture mechanics and the dotted lines to results based on data obtained by the Weibull method
for an effective surface of specimen of 1 cm2 (according to GoviW19!).

For ceramic materials, the Weibull parameter ranges from a few units to 20 or even
30: m = 6.4, 6.5 and 5.7 for α-SiC (Carborundum), at 25°C, 1200°C and 1475°C
respectively;^ m = 22 for a batch of RBSN Si3N4 at 25°C and m = 11, 19 and
20 for a batch Si3N4 + Si0 2 , at 20°C, 1000°C and 1100°Ο^ respectively.
Such statistical results can be very important and useful. For example,
GoviW19! tested 425 specimens of Si3N4 HP (Norton NC 132) at 20°C and 1371°C:
m = 6.5 and 7 respectively. This author obtained an excellent agreement with the
rupture stress calculated 1 cm 2 , a w , and the experimental value, σ^ corresponding
to a probability, V{ = 0.632, as can be seen in figure 2.1.7.
By use of other types of statistical tests, such as the "proof test" (survival
test), and by taking into account an eventual slow crack growth, it is possible to
determine the time to rupture and thereby the life prediction, or to estimate the
rupture stress and the structural reliability, i20-22] That only needs the knowledge
of n, A, στ and Kic, together with an a priori specification of the values of the
proof stress and the stress to be used· in service.
2.1.7 Time and temperature effects: creep behaviour
Ceramics generally exhibit good creep resistance as compared to metals,
but this depends on the microstructure. On a creep curve one generally observes a
1. Rupture and deformation of ceramics 41

primary creep which is dominated by viscoelastic deformation due to grain bound-


ary sliding accommodated by elastic deformation at grain boundary asperities
and/or adjacent grains, a secondary or stationary creep which is dominated by
diffusional creep and then a tertiary creep which is dominated by cavitation and
which is accentuated by subcritical growth of preexisting cracks S23^ In addition to
these main mechanisms, intragranular deformation through dislocation glide and
deformation due to a dissolution-precipitation mechanism can also arise.
Many mechanisms can be involved in creep damaging degradation, like
rupture with facets, intergranular fracture, cavitation, or exudation. Two of these
mechanisms are illustrated in figure 2.1.8J24,25^
Figure 2.1.9 displays respectively the creep deformation under compression
and the strain rate as a function of temperature for several ceramics J5'26^ One notes
the excellent creep resistance of these materials and the low values for the strain
rate.
Any in-depth interpretation of creep results requires diffusion data. Unfor-
tunately, very few diffusion investigations have been undertaken up-to-now. More-
over, one should also take into consideration the impurities and/or the secondary
phases, so that the problem appears, within the present state-of-the-art, as a really
ominous challenge.
2.1.8 Role of heterogeneities
We already indicated the role of the secondary phases at the triple points
or at the grain boundaries in downgrading the rupture characteristics, generally
by slow crack growth and grain boundary sliding, a mechanism which is facilitated
by the viscosity of the glassy phase.
For materials with a secondary phase, crack initiation very often originates
from this secondary phase. For materials without a secondary phase, fracture is
generally initiated from a defect: either a machined defect or a defect due to the
process method (inhomogeneity or impurity). That can be easily observed on scan-
ning electron micrographs.
We also have to recall that the critical size defect can be sometimes of the
order of magnitude of the mean ceramic grain size. In the extreme cases it can
even reach ten times that size. That explains why the initial ceramic powder and
the process parameters have to be controlled in the closest possible way.
2.1.9 Why high strength at high temperature?
A number of conditions must be met for ceramics to maintain high strength
at high temperature. Here are the main o n e s : ^
- strong covalent bonds, but this leads to a difficult sintering,
- density close to the theoretical value, at least for monolithic ceramics, but that
is probably not the case for ceramic matrix composites (CMC) where 10 or 20% of
porosity can be very useful,
- a regular and small mean grain size, which does not evolve with temperature in
42 J.-L. Chermant / Experimental evidences for various materials


/ 1

%
£ *>/ *f £''
A§>'BiRCH-1978

'
a.»
«y/
*x I
<''
1 1
/ *
1 i
t
t / '
11 0
/
// / /
1 i
1
1
1 s ^ ·
1 / / / //
1» ' // "'"' j ^
1 ' /
i / /
..···*' **' ' ' SiC
^*~ Si 3 N 4
1 " .·'
Sialon

Ψ/
0*

FN compression

V/^ 2J
CT :238MNrf
T Ι^δθ00
\ f
V
_J_L
100
1
_.
200 J__J

°C 1500 1400 1300


69MNm-2 HS 130
air

10-

\ (+7BeSiN 2 ) \SiAlON-Mg]
\
10- \
VNC400L\ \

S i C N C ,M
^ N C 2 l l \
HP+AI v

\
10-· c s o
Λ
Χ GPS 502
& ι
^ \(+7BeSiN2)

SI3N4 \

■X. SiAION-L.
_L·
5.5 5.7 5.9 6.1 V\°z*\tfm

Figure 2.1.8 Illustration of two types of mechanisms arising during creep of ceramics: (a) se-
quence formation of facets: crack initiation at a triple point, then after its opening by grain
boundary sliding (Evans^24!); (b) cavitation damaging: formation of a cavity at a triple point,
then extension by grain boundary sliding (Tsai and Rajl25^).
1. Rupture and deformation of ceramics 43

\__J
1 EVANS - 1U80
1 TSAI-1982 ^

^ f
~ \
* \
1 initial crack

t/i
Ht_ % /
~\
1 crack opening cavity initiation ^ 1

*
*
1 final crack growing microcrack 1

a RUPTURE WITH FACETS b. RUPTURE BYCAVITATION 1


' 1

Figure 2.1.9 Creep strain as function of time for different ceramics tested under compression
at 1350°C under 238 MPa (according to Birch and Wilshire[261) and strain rate, i, (under com-
pression for Si3N4 Norton NC 132) as a function of temperature for different ceramics tested in
3-point bending under 69 MPa.

the test range,


- perfectly controlled grain boundaries, since, as already seen, the presence of a
glassy phase in grain boundaries will decrease the strength characteristics in the
range of 600-800° C, a weakening which does not arise if grain boundaries are free
of secondary phase.
In conclusion, a completely reliable know-how has still not been established
for high-temperature/high-strength ceramics, but the characterization of the pro-
cess from the raw materials to the final product is without contest one of the most
fruitful lines of research. This is illustrated on a nice figure (fig. 2.1.10) due to
Ichinose,t27^ which summarizes the factors governing the strength of ceramics from
the point of view of microstructure, physical properties and external effects.

References
1. J. Clark and M. Flemings, Scientific American 255, 50 (1986)
2. British Ceramic Society, Ceram. Ind. J. 12, 3 (1979)
3. D.W. Richerson, Modern ceramic engineering: properties, processing, and use
in design, (Marcel Dekker, New York and Basel, 1982)
4. W.D. Kingery, H.K. Bowen and D.R. Uhlmann, Introduction to ceramics, 2nd
edition, (John Wiley and Sons, N.Y., 1976)
* Type of atom and its position

Figure 2.1.0 Factors controlling ceramic strength, from Ichinoset27! (with courtesy of N. Ichinose and John Wiley and Sons Ltd).
1. Rupture and deformation of ceramics 45

5. J.L. Chermant, Ceramiques thermomecaniques, (Les Presses du CNRS, Paris,


1989)
6. R.J. Brook, Mat. Sei. Eng. 71, 305 (1985)
7. M. Coster and J.L. Chermant, Precis d'analyse d'images, (Les Presses du
CNRS, Paris, 1989)
8. J. Hartley, Am. Ceram. Bull. 6 1 , 911 (1982)
9. H. Neuber, Kerbspanungslehre, (Springer Verlag, 1958)
10. F.E. Buresch, International Congress of Fracture, ICF 4, Waterloo, Canada,
ed. D.M.R. Tapplin, University of Waterloo, (Waterloo Press, 1977), Vol. 3, p.
929
11. F.E. Buresch, K. Frye and T. Müller, Fracture mechanics of ceramics, edited
by R.C. Bradt, A.G. Evans, D.P.H. Hasselman and F.F. Lange, (Plenum Press,
N.Y., 1983), Vol. 5, p. 591
12. S.M. Wiederhorn, Fracture mechanics of ceramics, edited by R.C. Bradt,
D.P.H. Hasselman and F.F. Lange, (Plenum Press, N.Y., 1974), Vol. 2, p. 613
13. A.G. Evans, S.M. Wiederhorn, Int. J. Fract. 10, 379 (1974)
14. W. Weibull, Ing. Veternkaps. Akad. Handlinger 151, 3 (1939)
15. W. Weibull, J. Appl. Mech. 8, 293 (1951)
16. A. de S. Jayatilaka, Fracture of engineering brittle materials, (Applied Science
Pub. Ltd, London, 1979)
17. S. Srinivasagopalan, M. Srinivasan and G.W. Weber, 41st Annual Conference
on Composites and advanced materials, (Coco-Beach, Florida, USA, 1980)
18. R.W. Davidge, G. Tappin and J.R. Mc Laren, Powd. Met. Int. 8, 110 (1976)
19. R.K. Govila, Am. Ceram. Bull. 62, 1251 (1983)
20. A.G. Evans and S.M. Wiederhorn, J. Mat. Sei. 9, 270 (1974)
21. S.M. Wiederhorn and N.J. Tighe, Proc. Workshop on Ceramics for advanced
heat engines, (Orlando, Florida, USA, Jan. 24-26, 1977)
22. S.M. Wiederhorn and E.R. Fuller, Mat. Sei. Eng. 71, 169 (1985)
23. F.F. Lange, B.I. Davis and D.R. Clarke, J. Mat. Sei. 15, 611 (1980)
24. A.G. Evans, Acta Met. 28, 1155 (1980)
25. A.L. Tsai and R. Raj, J. Am. Ceram. Soc. 63, 513 (1982)
26. J.M. Birch and B. Wilshire, J. Mat. Sei. 13, 2627 (1978)
27. N. Ichinose, Introduction to fine ceramics: applications in engineering, (John
Wiley and Sons Ltd, 1987)
46 D. Baptiste / Failure mechanismes of composite materials

2.2 Failure mechanisms of composite materials


Didier Baptiste*

2.2.1 Introduction
The use of composite materials is becoming more and more important
in industrial applications. The understanding of their rheological behaviour and
fracture is fundamental in order to predict the failure and its localization in a
structure. A composite material is obtained by mixing two or more components
having specific properties. The choice of the components is guided by the expected
characteristics of the studied structure. Reinforcements, which can be for example
long or short fibers, or particles of different sizes are introduced into a matrix.
In most cases, their mechanical properties - i.e. stiffness matrix, yield strength,
hardening parameters - have higher values than the one of the matrix. Depending
on volume fraction and orientation of the reinforcements, the matrix stiffness and
the failure strength can be considerably increased.
Damaging is a function of orientation, aspect ratio, local volume fraction
of the reinforcements, the mechanical properties of the components as well as the
characteristics of the reinforcement-matrix interface. All these parameters strongly
depend on the manufacture processes. Using identical initial components, the fab-
rication technique can lead to different local orientations, aspect ratios and volume
fractions of the reinforcements. In general these microstructural parameters scatter
considerably, inducing spatial fluctuations in the mechanical properties. The dam-
age process occurring inside the composite can also be different from one point to
another. For example, the damage does not propagate in the same direction if the
fibers are far apart or very close to each other. In this chapter only two families of
composites, namely materials with organic or metallic matrices, are studied.
2.2.2 Microscopic failure mechanisms of a unidirectional
fiberglass-reinforced composite
A fiberglass-reinforced composite, even if it is unidirectional, exhibits a vari-
ety of failure modes including: individual fiber fracture, interfacial debonding, ma-
trix cracking and delaminationJ 1- ^ We will investigate the tensile failure sequence
and corresponding mechanisms, especially at the scale of the fibers. Emphasis is
placed on understanding how the tensile behaviour is affected by manufacture flaws,
particularly the variation of local fiber volume fraction and fiber misaligment. We
believe that this particular study can be generalized to similar composites.
From the analysis by X-ray radiography and from the observations
through optical and scanning electron microscopes, it was found that when a uni-
directional specimen is gradually loaded in the longitudinal direction, the main
t Laboratoire MSS-MAT, U.R.A. CNRS 850, Ecole Centrale de Paris, F-92295 Chätenay-
Malabry Cedex, France.
2. Failure mechanisms of composite materials 47

Stress (MPa)
ii
1000- 1
ΙΠ
i Γ*Ί III: Fracture of a bundle of fibers
800-
jr
600- n
/
400- II: Individual fiber fracture

200- I I
1
1 1 1 1 1 I: Undamaged composite
0 1 2 3 4 Strain (%)

Figure 2.2.1 Tensile curve and schematic representation of the corresponding failure develop-
ment.

failure process consist of four stages:


(a) Individual fiber fractures at random locations inside the specimen. This begins
at less than 50% of the ultimate load, owing to flaws in the brittle fibers.
(b) Accumulation of fiber fractures in bundles of fibers at the specimen surface,
forming macroscopic cracks at different locations along this bundle.
(c) Initiation of delaminations parallel to the fibers, starting at the macrocracks.
(d) Propagation of delaminations in the fiber direction.
As the load further increases, failure continues by forming other cracks and
delamination along fiber bundles. These broken bundles then accumulate until the
specimen fails catastrophically.
A representative tensile curve (i.e. breaking characteristic) of the unidirec-
tional composite loaded in the longitudinal direction is shown in fig. 2.2.1. Three
regions can be distinguished in the curve:
• At less than 40 % of the ultimate stress, the curve is perfectly linear, which
evidently corresponds to the undamaged composite.
• Beyond this region, the tensile curve is no longer linear. In this case, the applied
stress is large enough to produce individual fiber fractures in the composite. The
slope of the curve decreases as the number of broken fibers increases, that is, as
the applied stress increases.
• At nearly 90% of the ultimate load, the tensile curve becomes highly non-linear
and the Young modulus of the composite decreases significantly. In this region, the
48 D. Baptiste / Failure mechanismes of composite materials

Figure 2.2.2 Failure surface of a unidirectional reinforced composite.

individual broken fibers accumulate to produce the fracture of bundles of fibers,


followed by delaminations parallel to the fibers at the specimen surfaces. The ul-
timate failure occurs when the remaining unbroken fiber bundles are unable to
sustain the applied load.
When the composite is subjected to a static tensile load in the longitudinal
direction, the fracture of an individual fiber within the composite creates two types
of stress concentration in its vicinity: one is the tensile stress concentration in the
first unbroken fibers, which may lead to their fracture; the other is the shear stress
concentration at the interface between the broken fiber and the surrounding matrix,
which tends to induce shear debonding along the fiber surface. Therefore, once a
fiber breaks, the local failure process can continue through two different possible
fracture modes: the tensile fracture in the adjacent fibers and the shear fracture at
the interface. Which fracture mode dominates directly depends on the local volume
fraction and the local fiber misorientation.
Usually, the fiber distribution in the composite is not homogeneous, and the
local fiber volume fraction can strongly deviate from the average. Such a variation
in fiber volume fraction affects the local failure process. Where the fibers are close
to each other, the fracture propagates from fiber to fiber in the tensile fracture
mode, and where the fibers are far apart, the failure progresses along individual
fiber surfaces in the shear fracture mode, giving rise to considerable fiber pull-out.

Fiber misalignment also influences the tensile strength of the composite.


Continuous, non-woven fibers in a composite, although unidirectional in principle,
display in fact small angular misalignment around their mean direction. The pres-
ence of angular misalignment, sometimes referred to as "fiber waviness", has been
2. Failure mechanisms of composite materials 49

Figure 2.2.3 Damage evolution during a tensile test (inside a S.E.M.) on an Al-SiCw composite.

recognized since many years and has been suspected of having a significant influ-
ence on properties such as longitudinal tensile modulus, longitudinal compression
strength, transverse mechanical properties and delamination fracture toughness.
The wider the fiber misalignment distribution, the smaller is the tensile strength.
2.2.3 Failure mechanisms in metal matrix composites
Metal matrix composites contain short fibers or particles which improve the
failure strength properties over that of the bare matrix. The damage mechanism
can be observed by testing a notched specimen under tension, inside a scanning
electron microscope (S.E.M.). In the case of an aluminium matrix with 20% silicon
carbide whiskers, the increase of the applied load gives rise to a failure process
which can be decomposed in the following steps (fig. 2.2.3):
• Failure of the interface at the tip of the fiber.
• Growth of a cavity at the fiber tip within the matrix.
• Coalescence of different cavities due to plastic instability. The path of the crack
depends on the relative position of the fibers. It may follow the interface between
the fibers and the matrix, or directly connect the cavities.
• Catastrophic propagation of the crack.
The failure, which is locally ductile inside the matrix, appears brittle on a
macroscale. The development of local plasticity around the fibers is correlated with
their relative positions. For a given volume fraction of fibers, the damage grows
starting from the tip perpendicular to the fiber, if the distance between two fibers
is small. This typically happens inside fiber bundles. For distant fibers, the damage
is still initiated at the tip, but evolves within the matrix, parallel to the fibers as
typically observed in the matrix around a bundle of fibers.
50 D. Baptiste / Failure mechanismes of composite materials

Stress (MPa) AS7GO3-10% SiC


A y, non-oxidized particles

200

AS7GO3-10% SiC
100 oxidized particles

Strain (%)
4 6 —10Mm

■10pm B ΙΟμπι

Figure 2.2.4 Stress-strain curves associated to the damage processes of two similar Al-SiC
composites with oxidized and non-oxidized particles. (Courtesy of R. da Silva^).

The aspect ratio, ί/d, of the reinforcement, (length divided by diameter)


considerably influences the value of the failure strength. For example, a composite
of a 2124 aluminium matrix with 20% of whiskers with an average aspect ratio of 5,
has a failure strength 25% higher than the composite with particles. The gain rises
up to 50% for an aspect ratio of 10. For higher values (ί/d > 50), the mechanical
properties reach a limit value. So very long fibers (ί/d > 100) do not increase the
failure strength further.
2. Failure mechanisms of composite materials 51

The properties of the interface between the reinforcement and the matrix
also has a very strong influence on the behaviour of the damage processes. Firstly,
the manufacture process produces internal stresses due to the different thermal
expansion coefficients of the matrix and the reinforcement. So the cooling of the
aluminium matrix from the melting temperature down to room temperature gen-
erates residual stresses. The value of these residual stresses depends on the tem-
perature. At 350° C, they completely vanish and the fracture surface of Al-SiCw
(where cw' stands for whiskers) displays fibers which were pulled out of the matrix.
No significant residues of the matrix are left on the fibers, confirming the fact that
the interface is weak. Secondly, chemical reactions take place at the interface. If
the interface is stronger, the composite will have a higher failure strength and a
lower ductility. In the case of an aluminium matrix composites reinforced with SiC
particles, the failure process is completely different if the particles are oxidized or
not. For the same AS7G03 matrix and reinforcement, the damage process will now
be compared for two different types of particle-matrix interface:
• The interface be strong because the particles are not oxidized: when the macro-
scopic plasticity starts, one observes first the fracture of the SiC particles. Then
when the load increases, the fragments of broken particles move apart, produc-
ing local stress concentration and shear bands in the matrix. Final failure occurs
through the propagation of these bands.
• The interface be weak because the particles are oxidized: a decohesion between
the particles and the matrix appears at the beginning of the macroscopic plasticity.
Cavities grow between the particle and the matrix, leading to stress concentrations
which are lower than in the previous case. Shear bands in the matrix are gener-
ated. An increase of the load intensifies these bands up to the final failure. Fig. 2.2.4
shows damage processes corresponding to the stress-strain curves of the two com-
posites. The grids which have a mesh-size of 3 μπι have been put on the specimen
surface, in order to measure the local strains.
2.2.4 Conclusion
The influence of different microstructural parameters on the rupture of
composites has been experimentally observed. The size of the reinforcements, the
distribution of their orientations, the local distance between them, the mechanical
properties of the components, the residual stresses and the quality of the interface
are the main parameters influencing the damage process.

References
1. L.A. Carlson and R.B. Pipes, Experimental Characterization of Advanced Com-
posite Materials, Prentice Hall, (1987)
2. M.R. Piggott, Load Bearing Fibre Composites, Pergamon Press, (1980)
3. K.L. Reifsnider, Damage in Composite Materials, A. S. T. M., (1980)
4. R. da Silva, D. Calmaison and T.Bretheau, Proceedings of MECAMAT 89,
Pergamon Press, (1989)
52 P. Acker / Concrete

2.3 Concrete: Large-scale heterogeneities and size effects


Paul Acker*

2.3.1 The structure of concrete


Since concrete consists of chemically inert inclusions of mineral origin,
called "aggregates" and a binder, cement, which hardens with water, it is fore-
most a composite material The mechanical properties (Young's moduli, strength
values) of the aggregates and the binder are often very different, and other physical
properties even more so.
The inclusions (sands and gravels) are characterized by a very large size
distribution (see fig. 2.3.1) covering more than two orders of magnitude (typically
from 0.1 to 20 mm) and therefore the concentration of inclusion is substantially
higher (of the order of 60%) than is generally found in what we normally call
"composite materials" (but which are less widely used than concretes). This sub-
stantially affects the analysis of the mechanical behaviour of this material, both at
the experimental level (in particular in the choice of scales of observation) and at
the theoretical level (limitations of homogenization methods, for example).

Porosity
length
scale

lnm Ιμπι lmm lm lkm
Cement
Granulometry Ultrafine^-^ FMerSand .
/ \ / \ Α Λ S— Graven

~i ' ' r~
lnm Ιμηη lmm
Structural E223 bridges
characteristic Ξ1 buildings
length ΕΞΞ3 structural elements
lab. samples
_1 r -^ ^_
lnm Ιμιτι lmm lm lkm

Cracks μcrack crack fracture

_1 ! ~Γ
lnm Ιμπι
Ιμηι lmm lm lkm

Figure 2.3.1 The various scales of heterogeneity of concrete.

t Laboratoire Central des Ponts et Chaussees, 58 Boulevard Lefebvre, F-75732 Paris Cedex
15, France.
3. Concrete: Large-scale heterogeneities and size effects 53

Because of the way it is made, (setting cement, a powder, in water), concrete


is a highly porous material. The total volume of pores is of the order of 10% (or
nearly a quarter of the volume of the matrix). For physical reasons (like the self-
desiccation, described below), the gaseous phase becomes continuous rather quickly
after setting. The pores, too, are characterized by a very large size distribution
(from nanometers to microns!) and by a "fractal" character (self-similarity over
several scales of observation): they have a highly ramified geometry, one which is
not at all like a network of cylindrical tubes but, on the contrary, one with a very
high level of connectivity among all the pore sizes and, in consequence, having
an enormous liquid/gas interface. This gives the material its highly hygroscopic
character (with no saturation in the moisture content) and makes it sensitive to
variations in temperature and humidity over the whole range of working conditions.
Within the material, three physical effects lead to large spatial variations
that cause internal stresses either on the scale of the inclusions or on the scale of
the concrete structuresJ 1 !
The first of these effects is of chemical origin. The volume of the hydrates
formed by the hydration of the cement is smaller than the sum of the volumes
of water and anhydrous cement needed. This shrinkage can be as large as 8 to
10% of the total volume. In fact, from the beginning, setting is determined by the
existence (and growth) of a continuous mineral skeleton. Since hydration continues
long after setting, under-pressure appears in the pores, accompanied by a phase
change and the appearance of a gaseous phase that quickly becomes continuous.
This is the phenomenon of "self-desiccation", which leads (in the matrix without
inclusions) to a homogeneous and isotropic shrinkage (endogenous shrinkage) which
is of the same order as the shrinkage that one would expect from the loss of water
due to the hydration process. In concrete, the apparent shrinkage is of course
much smaller (linear deformations of the order of 10~4 rather than 10 - 3 ) than
the potential "chemical" shrinkage, because the aggregates partially block this
deformation, producing non-negligible tensile stresses within the matrix.
The second effect is of thermal origin. The hydration reaction is highly
exothermal (200 to 400 joules per gram of cement), leading, under adiabatic condi-
tions, to a temperature rise that may exceed 50° C. This heat generation is generally
not relevant in laboratory specimens, but in structural elements it leads, during
the cooling period that follows setting, to temperature gradients and cracking of
the skin of the concrete. Moreover, since setting takes place at a higher temper-
ature, the cooling also leads to a compression of the aggregates (in addition to
that resulting from chemical shrinkage), since they often have a thermal expansion
coefficient smaller than that of the cement paste.
The last - and largest - effect results from the natural drying of the con-
crete. In spite of the self-desiccation described above, there always remains (except
in a few special concretes) a large quantity of free water (i.e. not chemically bonded)
in the pores of the binder (50 to 100 liters of water per m 3 of concrete, or more than
50% of the volume of the pores). This water will tend to migrate to the exterior,
54 P. Acker / Concrete

I Tangent modulus
II Rupture stress

30 -I
«3
PH PH
2H

CO 20 H
S
1H
10 H

2x10' :

Mean strain (Δ£/£) Mean strain (M/i)

Figure 2.3.2 The compressive strength is by definition the maximum stress reached (peak); the
decreasing part (post-peak) can be obtained only if the test specimen is subjected to an imposed
strain rate; the partial unloading cycles during loading show the limitations of the linear elastic
model. The tensile strength is substantially smaller than the compressive strength, and the cycles
applied before fracture exhibit a behaviour that is more nearly linear than in compression; even
when the test is at a controlled strain rate, it is exceptional to observe a post-peak behaviour.

more or less rapidly according to the climate. Even under extreme conditions, this
natural drying is very slow (ten years for a cube of 10~3 m 3 kept at 20°C and 50%
relative humidity!), and in large structures it remains a surface phenomenon, never
affecting a depth of more than 15 cm.
In massive structures (typically more than 1 m thick), the apparent
shrinkage observed during cooling or drying is only a secondary effect. Primar-
ily, one observes internal stresses (tensile at the skin, compressive in the core,
compressive at day joints) and cracks. The spacing between these cracks, and thus
their opening width, is directly related to the depth of cracking. During drying,
this mesh may vary from a millimeter to several meters. The opening width of the
skin cracks varies in the same range. It is this opening width that is of importance
to the engineer, on one hand for aesthetic reasons, but also because of durability:
above a certain critical opening width (which incidentally is close to the threshold
of visibility), the reinforcements are no longer protected from corrosion.

2.3.2 Rupture behaviour


Under simple compression, as under tension, the strength is defined as
the "peak" stress (see fig. 2.3.2), i.e. the maximum value reached during a test
conducted at constant imposed strain rate.
3. Concrete: Large-scale heterogeneities and size effects 55

Depending on the composition and the fabrication process of the concrete,


its compressive strength ac may range, under industrial conditions, from 20 to
60 MPa, but in special applications it may reach 120 or even 150 MPa. Its tensile
strength is more difficult to determine, and the experimental results are much
more dispersed (20 to 30%, against 5% in compression). It is much smaller, and
never exceeds 6 MPa. This is why engineers always strive to have concrete work in
compression, either through the design of the structure (arch, vault, prestressing
cables), or by reinforcing zones of tensile stress with metal reinforcements (which
take over the tensile forces, but only after the concrete has cracked).
The Young modulus of concrete ranges, from 25 to 50 GPa, and its Poisson
ratio is of the order of 0.17.
It has, in fact, no true elastic domain, such as found in metals. Even at low
stress levels, the emission of acoustic waves reveals the formation of microcracks
within the material, due to heterogeneities and, probably, to the existence of ini-
tial internal stresses. A large effort in research has made it possible, to characterize
this microcracking by its macroscopic effects: nonlinearities, damage (defined as
a change in elastic properties), and residual deformations. However, it has never
been possible to observe directly these internal microcracks, together with their
propagation mode, (i.e. in a way that would distinguish them from initial microc-
racks and from the microcracks generated by the preparation of the surface of the
sample).
The behaviour observed under loading cycles exhibits hysteresis (fig. 2.3.2),
which may be explained by friction between the crack lips, and residual strains,
related to the initial internal stresses.
The non-elastic strains under long-term loadings are very large (up to 2 or
even 3 times the instantaneous strains). In fact, a large part of these delayed strains
comes from a structural effect (cracking of the skin) during the transitional phase
of drying (which is very long, as we have seen). Under homogeneous conditions,
creep is proportional to the free water content, and this creep (basic creep) is, in
the long run (several years), close to the instantaneous strain.
2.3.3 Size effects
In normal applications, the heterogeneities caused by drying induce par-
ticularly large size effects. Not all consequences of this aspect have been fully
analyzed, it has therefore been impossible to exhibit the size effects of concrete
under compression.^
Under uniform conditions of temperature and relative humidity (already
rather difficult to achieve), such a size effect has been observed in tension (fig. 2.3.3).
This effect can be correlated quantitatively with the heterogeneities (fig. 2.3.4) by
means of a stochastic approach using the finite-element method:^ the peak in
the characteristic then corresponds to a transition from diffuse microcracking to
localized crackingJ4^
On the scale of a structure, the growth of a crack is accompanied by a
56 P. Acker / Concrete

5 0= 160 mm
Average = 2.52 MPa
0 jd|J r~Mli sdv = 0.08 MPa

8 5 0= 112 mm
!
fc
o
H , EÜ2L I ηπιη
Average = 2.75 MPa
sdv = 0.09 MPa

0 = 89 mm
5 Average = 2.96 MPa
jmja__EL sdv = 0.12 MPa
0 2.0 2.5 3.0 3.5 4.0 Strength
(MPa)

Figure 2.3.3 Influence of test specimen size on tensile strength of concrete. [3]

non-elastic zone that is often largeJ5^ This scale is therefore directly relevant to the
limits of applicability of linear fracture mechanics, in which the non-elastic zone
must be small with respect to the dimensions of the structure.
The failure of concrete under compression is a complex and still poorly
understood process, that is specially sensitive to boundary conditions.^ When
friction is eliminated on the surfaces where the stress is imposed, failure occurs in
small columns a few aggregates wide. The size effect seems to be less pronounced
than in tensile failure, probably because the regime of diffuse damaging is larger.
Concrete, which is neither plastic nor brittle, exhibits a particular mechan-
ical behaviour: because of its mineral character, it reacts to loadings by cracking;
because of its granular character, this cracking remains stable over a large domain,
allowing extensive use of this material in construction. But its use is still largely
based on empirical rules, and a better understanding of the mechanisms of fracture
would have considerable industrial impact.

References
1. P. Acker and M. Moranville-Regourd, Physico-Chemical Mechanisms of Con-
crete Cracking, in The Material Science of Concrete, chapt. 7, Am. Ceramic
Soc, (1990)
2. L. Elfgren ed., Fracture Mechanics of Concrete Structures. From theory to
applications, (RILEM Report), (Chapman and Hall, 1989)
3. P. Rossi and S. Richer, Numerical modelling of Concrete Cracking based on a
Stochastic Approach, Materials and Structures, 20, (1987)
4. J.M. Torrenti, J. Desrues, P. Acker and C. Boulay, Applications of stereophoto-
grammetry to the Strain Localization inConcrete Compression, in ref.[5]
3. Concrete: Large-scale heterogeneities and size effects 57

D = 5 cm
D= 10 cm
D= 20 cm
D= 40cm_
D= 60 cm

2 3 k
Strain χ 10u

Figure 2.3.4 Process of tensile failure of concrete, simulated by Rossi's stochastic model (perfect
brittle elastic contact elements with randomly distributed failure criterion values^).
58 P. Acker / Concrete

5. J. Mazars and Z.P. Bazant eds., Cracking and Damage : Strain Localization
and Size Effects, (Elsevier, 1989)
4. fracture mechanisms of metals 59

2.4 Fracture mechanisms of metals


Dominique Frangois*

2.4.1 Introduction
Most metals crystallize in simple systems, face centered cubic (FCC), body
centered cubic (BCC) or hexagonal. At the atomic scale, three types of fracture
can be considered (see fig. 2.4.1). One can find a separation of the atoms along a
crystallographic plane, either in a direction normal to that plane (mode I according
to fracture mechanics) or in a direction parallel to that plane (modes II and III)
or by the coalescence of vacancies, a mechanism which can occur at high enough
temperatures. The first type of separation is cleavage, while the second one is a
slip mechanism.

Figure 2.4.1 Schematic representation of the three types of cracking in metals on the atomic
scale. From left to right: cleavage along a dense crystallographic plane; slip along a dense crys-
tallographic plane; condensation of vacancies at high temperatures.

2.4.2 Cleavage
The separation of atoms requires a stress which is extremely high, of the
order of one tenth of Young's elastic modulus. This can only occur if some stress
enhancement mechanism takes place. In metals this is attributed to dislocations
piling up in front of obstacles such as grain boundaries or carbide inclusions. The
more localized the slip, the more dislocations pile up and the more likely becomes
cleavage. This explains why cleavage never occurs in FCC metals which have twelve
independent slip systems, providing efficient ways to relax the stress concentrations.
f
Laboratoire MSS-MAT, U.R.A. CNRS 850, Ecole Centrale de Paris, F-92295 Chätenay-
Malabry Cedex, France.
60 D. Frangois / Fracture mechanisms of metals

Figure 2.4.2 Cleavage fracture. On this fractograph taken in a scanning electron microscope,
facets corresponding to grains of different orientations are observed as well as river patterns
resulting from the merging of steps joining two cleavage planes which do not coincide. The
material is mild steel fractured by impact. The cleavage planes are {100} planes of BCC ferrite.

In BCC or hexagonal metals, cleavage is suppressed when slip spreads out at high
temperatures.
When the fracture surfaces of these metals, broken at low temperatures, are
observed with the naked eye or with a magnifying glass, a shiny crystallographic
appearance can be seen. In a scanning electron microscope (fig. 2.4.2) cleavage
fracture displays smooth surfaces with orientations which are determined by the
grains. It can be shown that they correspond indeed to typical crystallographic
planes, the cleavage planes, which are planes of high atomic density. Often typical
"river patterns" are observed on the cleavage plane. They consists of steps which
are due to small misorientations and which join as cleavage propagates, getting
higher and higher each time they meet. Thus these rivers "flow" in the direction
of the crack propagation.
Cleavage cracks are considered to be brittle at the microscopic level because
they involve little plastic deformation. Attention must be drawn however to the
fact that sometimes they are triggered only when strain-hardening has raised the
stress to a level, at which the macroscopic strain is not negligible.
Due to stress concentrations from localized slips, cleavage is more likely
when the grain size is large, when the temperature is low, or when the strain rate
is high.
4. Fracture mechanisms of metals 61

2.4.3 Ductile mechanisms


Single crystals of a hexagonal metal, such as zinc, can slip on a single plane
until complete separation of the two parts. More often, however, multiple slip occurs
in single crystals inducing a neck in the specimen loaded under tension. The section
of the crystal can then reduce drastically. In polycrystalline metals, necking takes
place in a more diffuse fashion, but can also lead to the complete separation of
the two halves of the sample when the cross-section of the neck reduces to zero.
This scenario is exceptional. In most cases, the specimen breaks sooner, and the
macroscopic appearance of the fractured pieces is that of "cup-and-cone" surfaces,
cup on one side, and cone on the other. This comes from the development in the
middle of the neck of a central crack perpendicular to the tensile axis, which at
the end tilts to a 45° orientation.
This central crack is the result of the coalescence of cavities which grew in
the central portion of the neck owing to plastic deformation. They can be observed
on microscopic sections taken of the neck. These necks usually initiate at inclusions
which do not deform in the same way as the metallic matrix. Stress concentrations
are thus created at the interfaces, producing either a cleavage crack inside the
brittle inclusion or fracture of the interface. Once these cracks are created, they
grow owing to the plastic deformation of the matrix and they become holes which
elongate in the direction of the maximum principal strain. The connecting regions
between them behave as small tensile specimens which at some stage neck down,
or are joined together by a localized shear instability. This coalescence of holes
coincides with the final fracture.
When observed with the naked eye or with a magnifying glass, the fracture
surface has, in this case, a fibrous appearance. In a scanning electron microscope,
dimples are observed (fig. 2.4.3). They are the coalesced cavities and the inclusions
where they were initiated can often be seen at one end of the dimples.
This fracture mechanism is ductile because it involves large local slip de-
formations and often it corresponds to a large macroscopic plastic deformation.
This is, however, not always the case and when the volume fraction of inclusions
is high, coalescence does not need large macroscopic deformation to occur. A clear
distinction must then be made between brittleness and ductility in the macroscopic
and the microscopic sense.
It is easy to understand that the hydrostatic part of the stress tensor,
i.e. its trace, has a strong influence on the growth of cavities. Hydrostatic pressure
prevents their growth, whereas a tensile positive hydrostatic stress increases it, thus
reducing greatly the fracture strain. The other important factor which influences
ductility is the volume fraction of inclusions, as was just explained.
62 D. Frangois / Fracture mechanisms of metals

Figure 2.4.3 Dimples resulting from the initiation at inclusions, growth and coalescence of holes
by plastic deformation. The inclusions can be seen at one end of the dimples. The material is
FCC austenitic stainless steel 316 LN.
4. Fracture mechanisms of metals 63

Figure 2.4.4 Intergranular fracture of steel. These fractures results from a weakening of grain
boundaries due for instance to a migration of impurities towards the surface of the grains.

2.4.4 Intergranular cracking


There are cases where the grain boundaries are weaker than the interior of
the grains. This is due to impurities gathering at the grain boundaries. If that is the
case, the cracks prefer to follow these grain boundaries resulting in intergranular
fracture (see fig. 2.4.4). In most cases, the fracture surface displays smooth facets.
But when impurities are gathered as precipitates or inclusions along the grain
boundaries, the result is dimple fracture.

2.4.5 Fatigue
Cyclic straining results in the progressive fracture of metals.
Generally, fatigue cracks starts at the surface. They are produced by ir-
reversible localized shear deformations, which, along what are called "persistent
slip bands", progressively roughen the surface on a microscopic scale. When this
roughening is strong enough, it becomes a crack which penetrates into the material
along the shear direction. In other cases, the surface slips hit the grain boundaries
which are then the sites of fatigue crack nucleation. An efficient way to prevent
this phenomenon is to produce compressive residual stresses at the surface, which
do not allow for the crack to open.
The small stage I surface fatique cracks meet obstacles when they propa-
64 D. Frangois / Fracture mechanisms of metals

gate towards the interior of the material, for instance grain boundaries, and they
then change their orientation to propagate in mode I perpendicular to the max-
imum principal stress. This is stage II crack propagation. Again it is the result
of irreversible plastic slips at the crack tip. At each cycle, the crack blunts and
propagates forward by a small distance. In many instances, this can be seen on the
fracture surface, in the scanning electron microscope, as striations (see fig. 2.4.5).

Figure 2.4.5 Fatigue fracture resulting from cyclic loading of austenitic, stainless steel 316 LN.
On this fractography, one sees striations which are left from local deformations along the fracture
tip at each cycle. These striations are however not always visible on fractographies of fatigue
fractures.

Corrosion often accelerates the fatigue cracking which is not simply the
cumulation of stress corrosion cracking and pure fatigue: synergetic mechanisms
take place making corrosion fatigue a dangerous phenomenon.

2.4.6 Creep cracking


When the temperature is high enough, vacancies can migrate by diffusion,
and they are likely to coalesce to create cavities and cracks. Two types of mecha-
nisms are observed in creep.
At relatively low temperature, and high stresses, slips along grain bound-
aries can create stress concentrations at triple points, and with the help of vacan-
cies, can open cracks along the grain boundaries which are perpendicular to the
maximum principal stress. This mechanism is called "W" (from wedge).
4. Fracture mechanisms of metals 65

At high temperature and low stresses various stress concentration mech-


anisms, such as dislocations, pile up on the grain boundaries. Also inclusion can
initiate holes which grow by the diffusion of vacancies and by plastic deformation.
They are all located along grain boundaries which are perpendicular to the max-
imum principal stress. Finally, the holes coalesce, producing a crack along these
boundaries. This cracking mechanism is called "R" (from round).
Fatigue and creep can interact: the grain boundaries weakened by cavitation
accelerate the fatigue crack propagation.
2.4.7 Conclusion
Cleavage, dimples, intergranular cracks, fatigue striations . . . a large variety
of morphologies are observed on fractographies of broken metallic pieces. The same
metal, according to the structural state induced by thermomechanical treatments,
to the state of stress and strain rate, to the temperature and to environments can
display various fracture mechanisms. With the help of the magnifying glass and of
the electron microscope, the metallurgist can read on fractographies a good part of
the life and of the death of samples. Mechanicians must be careful in their models
to take into account this large variability of mechanisms and to exercice care in
their extrapolations.

References
1. J.F. Knott, Fundamentals of Fracture Mechanics, (Butterworths, 1973)
2. D. Broek, Elementary Engineering Fracture Mechanics, (Noordhoff, 1974)
3. C.J. McMahon and M.Cohen, Acta Met. 13, 591 (1965)
4. J.F. Knott, J. Iron Steel Inst. 204, 104 (1966)
5. R.O. Ricchie, J.F. Knott and J.R. Rice, J. Mech. Phys. Solids, 21, 395 (1973)
6. D.A. Curry and J.F. Knott, Met. Science 10, 1 (1976)
7. D.A. Curry and J.F. Knott, Met. Science 13, 341 (1979)
8. A. Pineau, in Advances in Fracture Research, D. Frangois ed., (Pergamon, 1981)
p.553
66 L. Monnerie / Mechanical properties of polymeric materials

2.5 Mechanical properties of polymeric materials


Lucien Monnerie^

Polymeric materials are used more and more not only as commodity ma-
terials, e.g. in packaging, containers, and fibers, but also in technical applications
in aeronautics, and automotive industries. In these latter cases high technical per-
formance is required such as temperature resistance, high modulus and toughness.
In this chapter we will first briefly describe the various types of polymer
structures, and then we will present their mechanical properties, emphasizing the
responses specific to these materials. More details can be found in the references
cited in the text.
2.5.1 Polymer structures^1'2!
Over the years, polymer chemists have prepared materials with an enor-
mous variety of chemical structures. In addition there are numerous natural poly-
mers.
In many cases, the macromolecular chain consists of the repetition of a
single chemical unit (they are referred to as homopolymers). As an illustration the
chemical structures of the repeat unit of some of the most important polymeric
materials are shown in fig. 2.5.1. In some cases, in order to achieve a better com-
bination of properties, the polymer chain is made from a random distribution of
two (or more) chemical units (they are referred to as random copolymers). Some
important examples of these materials are the butadiene-styrene copolymers which
are extensively used in tires.
In addition to the type of units which are present (chain microstructure),
an important feature which controls the mechanical properties of polymers is the
chain macrostructure.
2.5.2 Macrostructures
The various types of polymer macrostructures are sketched in fig. 2.5.2.
• Linear chains. Many "thermoplastic" polymers like PS, PMMA, PA66, and
PETP have linear chains. The two characteristics of this macrostructure are :
- the polymer materials can usually be dissolved in an appropriate solvent
- at sufficiently high temperature, if no chemical degradation occurs, the materials
will be able to flow under a shear stress.
• Branched chains. Long chain branching can result either from the addition of
a small amount of a specific unit leading to the growth of a new polymer chain
on the backbone of a polymer, or from side-reactions. Branched polymers can be
dissolved and they flow but they exhibit melt rheology which is quite different from
that of linear polymers. In the solid state the mechanical properties of linear and
t Laboratoire de Physicochimie Structurale et Macromoleculaire, Ecole Superieure de Physique
et de Chimie Industrielles de Paris, 10 rue Vauquelin, F-75231 Paris Cedex 05, France.
5. Mechanical properties of polymeric materials 67

-CH2- -CH2-CH=CH-CH2- - C H 2 ~ C = CH~CH2- -CH2~CH-


CH3 CH
PE PB 3
PI PP

- CH2- CH- - CH2- CH- - C - - 0 - C - O - CH2- CH2- O -


2
| I II w II 2 2
c=o Ö ° °
O— CH3
3
PS PETP
PMMA
CH3
-^■C-^-O-C-O- -C-(CH2)4-C-NH-(CH2)6-NH-
CH3 O O O

PC PA66

Figure 2.5.1 Repeat units of some industrial polymers. Symbols are defined in table 2.5.1.

Linear Branched Network

Figure 2.5.2 Different polymer macrostructures.


68 L. Monnerie / Mechanical properties of polymeric materials

branched polymers are nearly identical.


• Polymer networks. In all "elastomeric" materials the polymer chains are cross-
linked by chemical agents in order to prevent the flow of the material under strain
and to obtain an elastic response. For elastomers, the cross-link density is typically
in the range of one cross-link every 100 to 200 repeat units. In contact with an
appropriate solvent the materials will swell but cannot be dissolved.
• Highly crosslinked networks. "Thermosetting" polymers, such as epoxy resins,
have this type of macrostructure. In this case, the chemical structure of the re-
actants is chosen in such a way that every one or two repeat units there will be
a cross-link point. Usually, such macromolecular structures cannot be swollen by
any solvent.
2.5.3 Molecular weight distribution
For linear and branched polymer chains, it is possible to define the molec-
ular weight (MW) of a chain as

Mi = tAfo, (2.5)

where M0 is the MW of the repeat unit and i is the number of units in the considered
chain. Polymeric materials, either natural or synthesized, are always composed of
chains with different molecular weights. This feature is characteristic of polymers.
The MW distribution is characterized through the number and weight av-
erage MWs and their ratio

V · N-M- Y- 7V-M·2 M~
Mn = Μ Ι= (2 6)
^vT' ^Ϋ~ΝΜ' π- ·
where Ni is the number of chains with i repeat units. To illustrate how large the
difference in chain MWs inside a polymer sample can be, three typical examples
of MW distribution curves are presented in fig. 2.5.3. The situation corresponding
to / = 1.05 is the narrowest MW distribution which can be achieved by anionic
polymerization in laboratory conditions. The case / = 2.1 is encountered in most
industrial polymers (PS, PMMA, PETP) where I values range from 1.9 to 2.6.
The largest MW distribution, / = 40, corresponds to industrial PE synthesized by
ionic coordinated polymerization.
It is easy to understand the great influence of MW distribution on physical
and mechanical properties of polymers by considering the example of polyethylene.
By increasing the number, n, of CH 2 repeat units in the chain one goes from
gas (1 < n < 5) to liquid (6 < n < 40), to grease (40 < n < 100), to wax
(100 < n < 500) and finally to a quite stiff solid for longer chains.
2.5.4 Crystallinity
Homopolymers with no substituents on their backbone, like PE, PETP
and PA66, are crystalline. However, owing to the length of the chains, polymeric
5. Mechanical properties of polymeric materials 69

1=1.05 1 = 2.1 1 = 40
N:4

4 5 3 4 5 3 4 5 6 7

Figure 2.5.3 Molecular weight distribution curves for three values of 7: 1.05, 2.1, and 40.

Amorphous
region

Crystalline
region

Figure 2.5.4 Chain organization in a crystalline polymer.

materials can never be entirely crystallized. There is always a mixture of crystalline


domains and amorphous regions, the chains participating in both. A schematic
organization of the chains in a crystalline polymer cooled from the melt is shown
in fig. 2.5.4. Depending on the cooling conditions, the crystallization extent lies
within the range of 20 to 55%.
In order for homopolymers with substituents, such as PP, PS and PMMA,
to crystallize they must possess a stereoregularity of the chemical structure (ίοτ
example all the substituents must lie on the same side of the plane defined by
the fully extended zig-zag conformation; such a stereostructure is called isotactic).
The industrial polymers PS and PMMA do not satisfy this requirement, and they
are entirely amorphous materials. Industrial PP is isotactic and consequently it is
crystalline.
It is worth noting that random copolymers are in general amorphous. This
is also the case for cross-linked polymers.
70 L. Monnerie / Mechanical properties of polymeric materials

fc
Έ ' 4
(Nm· 2 ) (Nnr 2 )

109

108 1(T

107
102

106
10

60 100 140 180 T (°C)

Figure 2.5.5 Temperature dependence of E' and E" for polystyrene — high MW; — low MW.

2.5.5 Temperature dependence of the physical state of polymers


Let us consider at first an amorphous polymeric material like PS. The
temperature dependence of the real and imaginary components (E1 and E" respec-
tively) of the complex modulus is shown in fig. 2.5.5. One observes a very large drop
of E' over a quite narrow temperature range around 100°C. Other measurements
like the specific volume, the thermal expansion coefficient, and the heat capacity
also exhibit an abrupt change in their temperature dependence in the same region.
The change from glassy, solid behaviour at low T to soft behaviour observed for
amorphous materials is called the glass transition, and the corresponding temper-
ature is denoted Tg.
At temperatures above Tg, the behaviour of amorphous polymers depends
on their MW. Polymer chains containing less than 200 bonds in their backbone are
liquid above T5, the melt viscosity being proportional to M w . For longer chains,
the entanglements existing among them prohibit the flow of the material. For
short time experiments (i.e. less than a minute) the material behaves as a rubber,
exhibiting a very high elasticity 1 . The modulus remains almost constant over a
temperature range which increases with the MW. At higher temperature, flow
behaviour gradually appears. In the flow regime the viscosity of long polymer
chains scales as M w ' .
The values of the glass transition temperature, Tg, (determined from dy-
namic mechanical measurements performed at 1 Hz) of some polymers are reported
1
Elastomeric materials are polymers used at temperatures higher than Tg + 50° C and which
are cross-linked to prevent flow and to make the highly elastic response permanent.
5. Mechanical properties of polymeric materials 71

Name Symbol T
(°C at 1 Hz) (°C)
P.Butadiene PB -90
P.Isoprene PI -60
P.Ethylene PE -100 +120
P.Propylene PP -20 +160
P.Styrene PS +105
P.Methyl-Methacrylate PMMA +115
P.Carbonate PC +140
P.Ethylene-Terephthalate PEPT +80 +210
P.Amide 6,6 PA66 +55 +260
Table 2.5.1 Glass transition temperature and melting temperature of some industrial polymers.

E k
(Nnr2)

109

107
-AV-
XN
-100 0 100 160 200 240 T (°C)

Figure 2.5.6 \og(E') versus temperature T for polyethylene (left), and for epoxy resins (right)
with different cross-link densities: — high; — moderate.

in table 2.5.1.
It has to be pointed out that the amorphous regions of crystalline polymers
also undergo a glass transition. However, due to the presence of crystalline domains
which retain their stiffness, the drop of E' is limited compared to that which occurs
for amorphous polymers. A further drop of E' occurs at the melting temperature
of the polymer (see T m values in table 2.5.1) as shown in fig. 2.5.6 (left).
For cross-linked polymers, the drop of E' at Tg decreases when the density
of cross-links is increased. Nevertheless, even for highly cross-linked epoxy resins, E'
decreases by a factor of 20 in the glass transition zone as shown in fig. 2.5.6 (right).
2.5.6 Mechanical properties of solid polymers^
The materials we will deal with in this paragraph are either amorphous
polymers below Tg or crystalline polymers below T m . The typical stress-strain
72 L. Monnerie / Mechanical properties of polymeric materials

.Yield point

\strain \Strain
softening hardening

Figure 2.5.7 Typical stress-strain curves of polymers.

PA66

£ 100

0.5 1.0 1.5


True strain
Figure 2.5.8 Examples of stress-strain curves.

curves in tension which are encountered are schematically drawn in fig. 2.5.7. Curve
A corresponds to brittle behaviour usually with a tensile strain at breaking of about
1-5%. Curve B exhibits a yield point and results in ductile fracture. In curve C,
the yield point is followed by the occurrence of necking which propagates along the
sample at almost constant stress. Strain hardening then leads to brittle fracture.
Curves obtained for various polymers at room temperature and at a strain rate
equal to 10~3 s - 1 are reported in fig. 2.5.8. We will consider each of these responses.

2.5.7 Yielding^
The yield point defines the onset of irreversible plastic deformation. It has
been shown to correspond to the maximum of the true stress in a compression test,
denoted σν. The value of ay is a function of the chemical structure of the material
and depends on the stress configuration (for yielding, it is mostly controlled by the
shear component of the applied stress). συ exhibits a logarithmic increase with the
5. Mechanical properties of polymeric materials 73

Crack I |T^I^

Figure 2.5.9 Cross-section of a polymer craze.

strain rate and a slow decrease with increasing temperature, leading to a vanishing
value at Tg. Applying a hydrostatic pressure results in an increase of ay.
2.5.8 Inhomogeneous deformation
Polymeric materials develop inhomogeneous deformations (necks, shear
bands) which are controlled by the geometrical constraints of the sample in a way
identical to other materials. In particular, under tension, the occurrence and devel-
opment of instabilities are well understood from the "Considere construction".^
2.5.9 Crazes^
Many solid polymers under tension exhibit between the elastic limit and
the yield point a series of crazes normal to the tensile stress. In transparent poly-
mers like PS or PMMA, such crazes are easily observed for they strongly scatter
visible light. Contrary to what happens for other materials, however, the inside of
a polymer craze is not empty but is filled in with polymer fibrils, the volume frac-
tion of which ranges from 0.1 to 0.5 (fig. 2.5.9). These fibrils are made of oriented
polymer (extracted from the lips of the craze) with a modulus much higher than
that of the unoriented polymeric material. Consequently, the overall modulus of a
sample after crazing is only slightly smaller than before, and the onset of crazing
cannot be noticed from the shape of the stress-strain curve. As shown in fig. 2.5.9,
the length of the fibrils gradually increases with the distance from the craze-tip.
At some critical length the fibrils break, leading to a crack which will later result
in a brittle fracture.
It is worthwhile to note that crazes occur both in amorphous and crystalline
polymers with the same characteristics.
2.5.10 Shear bands^
Polymers are able to develop shear bands both under tension and com-
pression. Depending on the chemical structure of the polymer, the temperature,
and the strain rate, έ, the shear bands present different structures as illustrated in
fig. 2.5.10. The shear bands are diffuse at high T or low έ and become localized
at lower T or higher έ. Upon further deformation the diffuse shear bands lead to
ductile fracture, whereas the localized shear bands end up in brittle fracture.
74 L. Monnerie / Mechanical properties of polymeric materials

Figure 2.5.10 Types of shear bands observed between cross-polarizers.

The crossing of two shear bands quite frequently initiates a craze. The
stress concentration which develops at the craze-tip can initiate shear bands as
well. Thus, under experimental conditions where shear bands and crazes are both
possible, their interaction can significantly affect the ductile-brittle behaviour of a
polymeric material.
2.5.11 Criteria for plastic deformation
The classical criteria (Tresca, von Mises) used for yielding in other mate-
rials can be satisfactorily applied to polymers. A pressure dependence has to be
considered, however, leading to the following modified expression for the von Mises
criterion

y ^ - <T2)2 + (σ2 - σΆγ + (σ3 - σι)


2
= τ0(έ,Τ) - μ(έ,Τ)ρ, (2.7)

where σ^ are the principal stresses, and r 0 and μ are characteristics of the polymer
materials which depend on έ and T.
A specific criterion has been proposed for crazing of polymers, based on a
critical tensile strain :

In this expression E is the Young modulus, and (7(ε,Τ) and D(i,T) are functions
characteristic for individual polymers.
In fig. 2.5.11, the representation of these criteria for PMMA at 20°C under
biaxial stresses is shown.
2.5.12 Brittle-ductile transition
The brittle strength of polymeric materials has a weak temperature de-
pendence as well as a small strain rate sensitivity. On the other hand, these two
5. Mechanical properties of polymeric materials 75

F i g u r e 2.5.11 Yielding and crazing criteria for PMMA in biaxial deformation.

σ
Brittle
stress
High έ
Low έ
Yield
stress

T
Figure 2.5.12 Temperature and strain rate dependence of brittle stress and yield stress.

factors strongly affect the yield stress value, as shown in fig. 2.5.12. This results in
a transition from brittle to ductile fracture as a function of temperature. Such a
transition occurs for PMMA in the temperature range 0 - 45°C and around 90°C
for PS.

2.5.13 Toughening of polymers


For many technical applications, most solid polymers, except polycarbon-
ate, do not have a sufficiently high impact resistance . The same is true for highly
cross-linked polymers like epoxy resins.
In order to improve the toughness of these materials, elastomer particles of
diameters in the range 0.1 to 3 μπι, can be dispersed in the polymer matrix. For
brittle matrices (PS for example) where crazing is the dominant failure factor an
increase of toughness is obtained by the following mechanisms :
- owing to the lower modulus of elastomer particles as compared to that of the
matrix, a tensile stress intensification occurs at the equator of each spherical par-
ticle. This leads to the onset of crazes at a lower value of the applied stress and
76 L. Monnerie / Mechanical properties of polymeric materials

consequently to the development of a larger number of crazes in the front of a


propagating crack,
- the elastomer particles are able to stop the growth of crazes and thus to avoid
the transformation of most of them into catastrophic cracks.
The development of many crazes in a brittle material in front of a propa-
gating crack absorbs an important amount of energy. As an example, the impact
strength of PS at 25°C is increased from 0.2 to 1.2 J/cm by the addition of 0.2
volume fraction of polybutadiene particles with diameters of around 3 μπι.
In the case of more ductile polymer matrices the best choice for the size
of rubber particles is around 0.2 //m. Indeed, the interaction between crazes and
shear bands contributes to stop the craze growth and avoid the formation of catas-
trophic cracks. PMMA and PVC belong to this class of matrices, and their impact
properties are improved in this way for industrial applications.
2.5.14 Conclusion
In this section, the main features of polymeric materials have been sum-
marized with a particular emphasis on mechanical properties of solid polymers.
It is worth pointing out that the use of polymers in technical applications
is increasing very rapidly. The automotive industry has become one of the largest
consumers of technological "plastics" not only for decorative purpose inside cars
but now as structural elements as well.
In spite of the present understanding of properties, there is still need for
research to determine the relationship between the chemical and physical structure,
and the mechanical properties of polymers.

References
1. G. Champetier and L. Monnerie, Introduction a la Chimie Macromoleculaire,
(Masson, Paris, 1969)
2. F.A. Bovey and F.H. Winslow, Macromolecules : An Introduction to Polymer
Science, (Academic Press, New York, 1979)
3. I. Ward, Mechanical Properties of Solid Polymers, 2nd ed., (John Wiley, New
York, 1983)
4. A.D. Jenkins ed., Polymer Science, Vol. 1, (North Holland Publ., Amsterdam,
1972)
5. H.H. Kausch ed., Crazing in Polymers, Adv. Polym. Sei., 52/53, (Springer-
Verlag, Berlin, 1983)
6. R.N. Haward ed., The Physics of Glassy Polymers, (Applied Science Publ.,
London, 1973)
6. Viscous fingering and viscoelastic fracture in clays 77

2.6 Viscous fingering and viscoelastic fracture in clays


Henri Van Damme and Emmanuel Lemaire*

2.6.1 Clays, clay-based fluids and clay-based solids^1,2]


Clay minerals are ubiquitous in nature. They occur in discrete geological
deposits, in soils and in all products of erosion and breakdown of rocks. They are,
by definition, the fine-grained (sizes below about one micron) fraction of minerals.
It turns out that this purely granulometric definition closely corresponds to a par-
ticular class of the silicates with a layer lattice structure: the "phyllo silicates", of
which micas are probably the most widely known.
The essential feature of phyllosilicates is their two-dimensional atomic
structure, which is often reflected by the platy shape of their microcrystals. The
basic structural units are layers of oxygen ions, which are obtained by linking a
tetrahedral sheet to an octahedral sheet (1:1 layers, « 0.7 nm thick) or by sand-
wiching an octahedral sheet between two tetrahedral sheets (2:1 layers, « 1 nm
thick). The 1:1 and 2:1 layers are not necessarily electrically neutral, due to unbal-
anced ionic substitutions in the cationic lattice which lead to a net negative charge.
The total charge balance is maintained by interlayer species, which may be either
individual cations as in micas or hydrated cations as in "smectites". The layer
type, the charge on the layer and the nature of the interlayer material provides the
basis for the classification of clay minerals. Each of them displays very distinct and
typical structural, morphological, colloidal and surface-chemical properties.
The most interesting family of clays, as far as physical properties are con-
cerned, is the family of smectites. Their 2:1 layer charge is large and is compensated
by hydrated interlayer cations which, in natural clays, are essentially hydrated
sodium, calcium or magnesium ions. The hydration energy of the interlayer ions
allows water to enter the interlayer space. This makes the original interlayer cations
easily exchangeable by other metal cations or by any other cationic species. Each
2:1 layer has an extremely high aspect ratio. Its lateral extension is of the order of
500 nm and it is only 1 nm thick.
Like other clays and other colloids, smectites can appear in a wide range of
physical states. When dispersed as dilute suspensions in simple liquids, they form
weakly non-Newtonian fluids (muds). At higher concentration, larger departure
from Newtonian behaviour is observed and they often form viscoelastic, shear-
thinning pastes. At still higher concentration they can form physical gels. Finally,
when all the dispersion liquid is removed or expelled for instance by evaporation
or compaction, they form solids (rocks), which can spontaneously crack under the
action of the stresses developed by the removal of water (figure 2.6.1).
t C.N.R.S.-C.R.S.O.C.L, F-45071 Orleans Cedex 02, France.
78 H. van Damme and E. Lemaire / Viscous fingering and viscoelastic fracture

Figure 2.6.1 Cracking pattern of a clay-rich mud after drying.

Among the best known properties of smectites is their ability to swell and
to shrink upon addition or removal of water. For that reason, smectites are often
referred to as swelling clays. The swelling stems from hydration of the interlayer
cations, from osmotic forces and probably also from shape fluctuations. Swelling
is associated with a statistical increase of the average distance between layers.^
However, even in dilute suspensions (soils), the layers are not totally isolated.^ The
basic units in a slurry of natural sodium-smectite are aggregates of a few layers,
more or less parallel to each other. At higher concentration, the aggregates start
forming a connected solid network (figure 2.6.2) with a lenticular porosity.^
The size of the aggregates (that is, the average number of stacked layers)
strongly depends on the nature of the interlayer ionsJ5^ The general trend is that,
the higher the charge of the cation, the larger the number of stacked layers. This
is clearly due to short range attractive forces between the interlayer cations, which
act as sticking points, and the oxygen ions of the layer surface. The increase of the
number of stacked layers has a direct micromechanical consequence: it stiffens the
particles. Thus depending on the interlayer ions, the morpho-mechanical properties
of smectite particles can go from highly flexible and thin sheets to thicker, more
rigid platelets.
6. Viscous ßngering and viscoelastic fracture in clays 79

Figure 2.6.2 Digitized transmission electron micrograph of a water-based smectite mud (by
courtesy of D. Tessier).

2.6.2 Viscous fingering


In simple shear flow smectite-based muds often display shear-thinning be-
haviour with a yield stress, which is described by the Herschel-Bukley equation:^

σ5 = σ γ + / ? 7 η , (2.9)

where σ5 = \σνχ\ (ayx is the stress tensor defined in chapter 3), 7 = \iyx\ (iyx is the
strain rate tensor), σγ is the yield stress and n < 1 is the shear-thinning exponent.
β has dimensions Pa · s n . Thus the non-Newtonian viscosity, η = σ δ /7, decreases
as 7 increases. For n = 1, eq. (1) describes a Bingham fluid, whereas for σ γ = 0
and n < 1, it describes a simple shear-thinning fluid.
In addition, concentrated muds often display viscoelastic properties, with
an elastic character which increasingly shows up as the characteristic time of the
flow event, TA, approaches (from above) the time taken by the microscopic reor-
ganization processes in the fluid, r str ("structural relaxation time"). This can be
quantified using the "Deborah number"^

De = r s t r /r f l . (2.10)

For a characteristic experimental length / and a velocity v, TA = I/v. The structural


relaxation time is less obvious and is related to the reciprocal frequencies of the
colloidal particles motion in the fluid. For a simple estimate one can use Maxwell's
80 if. van Damme and E. Lemaire / Viscous ßngering and viscoelastic fracture

relaxation time, r str = μ/G, for a body characterized by a Newtonian viscosity, μ,


and an elastic modulus, G. De may be interpreted as the ratio of elastic to viscous
forces. For D e C l , viscous effects dominate, whereas for De > 1, the system will
essentially behave as an elastic solid.
Since they possess both non-Newtonian and viscoelastic properties, muds
may exhibit extremely unstable behaviour. This is most easily studied by looking
at the intrusion of a simple low viscosity fluid under pressure into a mud, within a
Hele-Shaw cell (two glass plates, separated by a narrow gap). This either induces
"viscous fingering", when the invaded medium responds like a fluid, or "hydraulic
fracturing", when the invaded medium responds like an elastic solid.
Viscous fingering is the interfacial instability phenomenon one refers to
when a high-viscosity fluid is displaced by a low-viscosity fluid within a medium
where flow is dominated by viscous friction (see chapter 8 for more details). The
shape of the interface is controlled by the balance between the destabilizing ac-
tion of the viscous forces and the stabilizing action of the interfacial tension (for
immiscible fluids) or interfacial dispersion (for miscible fluids). A suitable control
parameter is the capillary number, Ca, or its modified form, Ca', in the case of
Hele-Shaw channels:
Ca = μυ/Γ and Ca' = 12(w/b)2Ca. (2.11)
v is the velocity of the interface, μ the viscosity of the displaced fluid, Γ the
interfacial tension, b the distance between the two plates of the cell, and w the
width of the channel.
Viscous fingering exhibits two extreme morphologies in Hele-Shaw chan-
nels. At low capillary number it leads, with immiscible Newtonian fluids, to smooth
"Saffman-Taylor fingers".^ On the other hand, when the displaced fluid is non-
Newtonian, it leads extremely easily to highly branched fractal patternsJ 10-13 ^
When the effective capillary number Ca' = Ι2(ιν/Β)2(ην/Γ) is calculated is these
experiment,^ one notices that extremely high values are reached, of the order of
106.
The morphogenetic process which turns a smooth Saffman-Taylor finger
into a highly branched fractal pattern can easily be analyzed with muds, by in-
creasing either the injection pressure or the clay particle concentration in the mud,
or both. This is shown in figure 2.6.3. It reveals various instability modes, includ-
ing the "tip-wobbling" mode, the asymmetric "hump" mode, and tip-splitting. The
sequence was o b t a i n e d ^ by injecting air into weakly non-Newtonian muds (water
with a weight fraction of 3% of clay; patterns 1 to 5), and then in muds of increasing
concentration (from 5 to 9%; patterns 6 to 10). This raises the yield stress from 0.4
to 137 Pa and decreases the shear-thinning exponent from 0.73 to 0.45. The range
for the effective capillary number, Ca', covered by the sequence runs from 2 x 103
to 7 x 106. A detailed examination of the sequence shows that the development
of the mass fractal morphology occurs via two parallel processes: first through the
development of a fractal interface, due to the increasing gap between the average
6. Viscous fingering and viscoelastic fracture in clays 81

2.6 x10 3 1.3 χ10 4 2.6χ104 3.9χ104 55x10 4

4x106 4.5x106 5x106 6x106 7x106

Figure 2.6.3 Viscous fingering patterns obtained by injecting air into aqueous clay muds at
increasing capillary control parameter, Co!.
82 H. van Damme and E. Lemaire / Viscous ßngering and viscoelastic fracture

finger or branch width and the pattern width (two lengths which merge in Saffman-
Taylor fingers), and, second, through the selection of a well-defined branch width,
which turns the surface fractal morphology into a mass fractal morphology.
2.6.3 Viscoelastic fracturing
Viscous fingering experiments performed at high velocity in concentrated
muds lead to significant modifications of the tip profile and the branching angle.
The tip profile, which was essentially convex, becomes concave and the branching
angle, which was rather closed ( « 30°) opens to almost perpendicular branching. ^
Quasi-perpendicular branching and concave tips in concentrated pastes
were taken as precursory signs for the onset of fracture by Van Damme et alM2^
They suggested that if a finger tip grows at a velocity larger than //r s t r , a stress
would develop behind the tip, along the finger. This stress would be oriented par-
allel to the finger and would grow as the finger moves forward. At some point, a
lateral crack could grow to relax the stress, and the orientation of this crack would
be naturally at 90° with respect to the main finger. As the tip keeps moving, the
lateral stress could grow again and a second lateral crack could form, and so on,
leading, eventually, to an apparently periodic lateral branching.
The clear confirmation of this came when even more concentrated pastes
(solid/liquid ratio between 0.1 and 0.6 weight fractions) were studied in radial
geometry^15! The general morphology of the patterns was found to be much closer
to what one intuitively expects for fracture. The growth sequences, like in figure
2.6.4 for instance, clearly show that bifurcation may indeed take place behind the
tip. The cracks are generally bent and the subcracks systematically grow on the
convex side of the main crack, at 90° with respect to its tangent. Growth also takes
place through a widening of the cracks, essentially on their convex side.
The strong morphological differences between the fingering patterns and the
fracturing patterns permit, in most cases, to identify a given pattern unambiguously
in either one of these categories. This in turn allows one to construct a fingering-
fracturing "phase" diagram in the plane spanned by the injection flow rate and the
colloid concentration in the paste (which is indirectly and non-linearly related to the
cohesion energy and the elastic properties of the paste). As shown in figure 2.6.5, the
transition between the two regimes is roughly hyperbolic, which is in agreement
with the idea that the crossover occurs when the frequency of the mechanical
excitation exceeds the frequency of the internal relaxation processes (the stronger
the paste, the smaller the frequency of internal microscopic reorganizations and
the smaller the flow rate necessary to exceed that frequency). A crude e s t i m a t e ^
shows the De is close to unity at the onset of fracturing.
The fracturing regime which is revealed by these experiments still raises
many questions, for instance on the components of the stress tensor. It cannot
be compared directly to type I crack opening nor to three-dimensional hydraulic
fracturing, where one expects crack bifurcation to occur at the tip, with an angle
smaller than 90°. A major difference is that friction with the cell walls is still an
6. Viscous ßngering and viscoelastic fracture in clays 83

Figure 2.6.4 Growth sequence of fracture patterns. Note the growth of cracks behind the tip.
84 H. van Damme and E. Lemaire / Viscous nngering and viscoelastic fracture

D
°! I
T «
ΐ i■ a 4 D

E
\

- M"
^
ui 10 F ■
H \
< *·, \X
cc
5
VF\
§- I □s^B %
• ·\
u. •
cm ·
1 tot Ä i >·-·^··· ^ - T I I I I ^ I

0.1 a.3 0.5 S/L [vy/w]

Figure 2.6.5 Morphologic "phase" diagram in the plane spanned by the injection flow rate
and the colloid concentration. Circles and squares correspond to viscous fingering and fracturing
patterns, respectively. Open and closed symbols correspond to miscible and immiscible intru-
sion, respectively. The curves are the estimated boundaries between viscous fingering (VF) and
fracturing (F), in the miscible (dotted curve) and immiscible (continuous curve) case.

important parameter here.

References
1 H. van Olphen, Introduction to Clay Colloid Chemistry, (Wiley, New York,
1963)
2. Clay Minerals: Their Structure, Behaviour and Use, G.Brown ed., (The Miner-
alogical Society, London, 1984)
3. K. Norrish, Disc. Faraday Soc. 18, 120 (1954)
4. J.J. Fripiat, J. Cases, M. Frangois and M. Letellier, J. Coll. Interf. Sc. 89, 378
(1982)
5. H. Ben Rhaiem, C.H. Pons and D. Tessier, Proc. Int. Clay Conf., Denver 1985,
L.G. Schultz, H. Van Olphen and F.A. Mumpton eds., (The Clay Minerals
Society, Bloomington, p.292, 1987)
6. R.B. Bird, R.C. Armstong and 0 . Hassager, Dynamics of Polymeric Liquids,
Vol.1, (Wiley, New York, 1987)
7. G.M. Homsy, Ann. Rev. Fluid Mech. 19, 271 (1987)
8. D. Mader, Hydraulic Proppant Fracturing and Gravel Packing, (Elsevier, Am-
sterdam, 1989)
9. P.G. Saffman and G.I. Taylor, Proc. R. Soc. London A 245, 312 (1958)
10. J. Nittmann, G. Daccord and H.E. Stanley, Nature 314, 141 (1985)
H . H . Van Damme, C. Laroche and L. Gatineau, Rev. Phys. Appl. 22, 241 (1987)
6. Viscous ßngering and viscoelastic fracture in clays 85

12. H. Van Damme, C. Laroche, L. Gatineau and P. Levitz, J. Physique 48, 1121
(1987)
13. H. Van Damme, E. Alsac, C. Laroche and L. Gatineau, Europhys. Lett. 5, 25
(1988)
14. H. Van Damme, E. Alsac et C. Laroche, C. R. Acad. Sc, I I 309, 11 (1989)
15. E. Lemaire and H. Van Damme, C. R. Acad. S c , II 309, 859 (1989) (1989)
16. H. Van Damme, in The Fractal Approach to Heterogeneous Chemistry, D.
Avnir ed., p.199, (Wiley, 1989)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
©Elsevier Science Publishers B.V. (North-Holland), 1990 87

3. Continuum and discrete description of elasticity and


other rheological behaviour

Stephane Roux*

In the following, we propose a very short introduction to linear elasticity,


and other simple rheological laws. We will mainly focus on the properties that
should be taken into account if one wants to make a realistic discretization of the
basic equations. We insist in particular on some models from last century used in
the framework of strength of materials, but which are not frequently encountered
in physics textbooks. Obviously, many important points are left aside, and the
reader is referred to classical textbooks of linear elasticity, rheology of solids and
strength of materials for further information. Somes general references are given
at the end of this chapter.
This chapter is organized as follows: We first give (section 3.1) some notions
of elasticity in the three-dimensional continuum formulation. In order to show that
this formulation is the most simple linear theory one can consider, we compare it
to electrical conductivity. Next we turn (section 3.2) to the classical modelization
of more complex rheological behaviours. We will see that if viscoelasticity can be
considered as a simple extension of elasticity, at least formally, this is not the
case for plasticity. In particular plasticity which includes work hardening requires
a relatively heavy formalism. We mention some directions for the modelization of
damage, as a rheological behaviour. The following section (3.3) comes back on much
more simple concepts, i.e. the one-dimensional versions of the three-dimensional
rheological laws. These constitute the elementary bricks with which we will build
networks. This approach to a discrete description of a two- or three-dimensional
solid is an alternative to the one exposed in the following section (3.4), which
is more mathematically founded. These two approaches to the discretization of
continuum solid are finally compared.
t
LPMMH, U.R.A. CNRS 857, Ecole Superieure de Physique et Chimie Industrielles de Paris,
F-75231 Paris Cedex 05, France.
88 S. Roux / Continuum and discrete description of elasticity

3.1 Elasticity
3.1.1 Linearity
We begin with the basic notions used in the construction of linear elasticity
and we will refer as much as possible to the corresponding concepts used for linear
conductivity.
As a very first rule, we impose that the theory we will build should be
invariant with respect to any Galilean transformation: Any translation or rotation
or combination of both, added to the displacement field of a medium should neither
change its energy, nor any physical quantity we can derive from it such as the
stress field. This requirement is most naturally expressed in terms of displacements.
However, it can also be formulated when one deals with stresses, in the form of a
balance equation.
Moreover, we require that the theory is linear. This is an approximation
that is only valid for small strains. This choice limits the range of applicability of
the theory as we will see below.
The displacement field inside the solid will be characterized by a vector
field U(x). In principle, one should distinguish between Eulerian and Lagrangian
coordinates, that is to say, we should now specify whether the x we refer to in
this displacement field is the undeformed original medium, or the actual medium
which has already been deformed. However, since we will only introduce here linear
elasticity, we can use one or the other indifferently. Let us consider that all defor-
mations are considered with respect to the undeformed geometry. The deformation
of the reference medium is for instance also encountered in fluid mechanics, and it
gives rise to the non-linear convective derivative in the Navier-Stokes equation. The
situation is very similar here. However, it is very usual that, in practical situations,
elastic solids are subjected to small strains and we will keep this hypothesis in the
following in order to build a linear theory. The above non-linearity is referred to in
the following as "large strain non-linearity".
We also mention now another source of non-linearities which arises from the
change in the original geometry: contacts. Suppose, that a crack exists in a medium.
Then depending on the direction of the force applied to the system, the borders of
the crack can either come into contact, or move apart. Obviously, any satisfactory
description of such a medium will require a different description depending on the
sign of a force. This is clearly out of reach of any linear theory. Therefore, from now
on, we will consider that the geometry does not evolve under the applied stresses,
unless explicitly mentioned.
There are other sources of non-linearity. A very general one is the depar-
ture from proportionality between the stress and the strain. These effect will be
called rheological non-linearity. We will give some examples for these non-linear
rheological laws, such as rupture or plasticity.
The equivalent of the displacement field in a conductivity problem is the
potential field. In this respect, elasticity is sometimes termed a "vector" problem as
1. Elasticity 89

opposed to the "scalar" case of conductivity. The Galilean invariance now translates
into the invariance of the physical properties under the addition of a curl field to
the potential. The large-strain non-linearities have no equivalent in conductivity,
but the rheological ones appear as a non-proportionality between the current and
the electric field (non-ohmic conductors).
3.1.2 Linear elasticity^
We choose to characterize the elastic deformations of a medium by the
displacement field U(x). To comply with the requirement of translational invari-
ance, we express the elastic energy of a solid as a function of the derivatives of
the displacement field. Since the displacement is a vector field, the gradient is a
second order tensor whose component α,/3 can be written daUß. We can split this
gradient into a symmetric and an antisymmetric tensor. The latter will contain
the rotational part of the field. Because of invariance under rotations, we require
that the energy should not depend on the antisymmetric part of the gradient of
displacement. We thus call "strain field" the symmetric part of the gradient of the
displacement field, εαβ,
eaß = Ud«Uß + dßUa). (3.1)
To simplify the discussion, we suppose that we are dealing with a homogeneous and
isotropic solid. In order to have an intrinsic form for the energy, i.e. which does not
depend on the basis chosen to express the displacement field, we need to introduce
only invariants of the tensor ε. A second order tensor in three dimensions, has only
three invariants:

ε\ = εαα = tr(e),
£π = εαβεβα = tr(e 2 ), (3.2)
£m = εαβ€β7ε7α = ίτ(ε ).
In order to build a linear theory, we introduce an expression which is quadratic
in the strain field. Thus only the square of the first invariant, and the second
invariant can be used and we are naturally led to the general expression of the
elastic energy £:
ε = \ (2//tr(£ 2 ) + Atr(s) 2 ) , (3.3)
where λ and μ are material-dependent constants, which characterize the rigidity
of solids. They are called Lame's coefficients.
We now introduce the conjugate field to the strain, called the "stress field".
By definition, this stress field, which is also a second-order symmetric tensor field,
σα/3, is related to the the strain field by

<?αβ = ~ , (3.4)
όεαβ
or
σ<*β = 2μεαβ + \δαβ{εΊΊ) (3.5)
90 S. Roux / Continuum and discrete description of elasticity

Figure 3.1 The resulting force F exerted on the surface dS of an infinitesimal volume element
is given by the product of the stress tensor with the vector n normal to the surface.

where δ is Kronecker's symbol. The linearity of equation (3.5), called Hooke's law,
is nothing but the consequence of our previous choice for a linear theory.
The conjugate variable of a displacement is a force (the product of both
being an energy). We therefore expect a relation between the stress tensor and the
force distribution inside a solid. It is indeed easy to work out for a uniform strain
in a cubic solid, the force that has to be applied on one face to equilibrate the
stress. The result is shown in figure 3.1. For an infinitesimal surface d 5 , where the
stress field can be considered homogeneous, the force F is given by

Fa = σα0ηβ dS, (3.6)

where n is a unit vector normal to the surface dS. One should also note that due
to the symmetry of the stress tensor, there is no local torque. We recall that the
symmetry of σ comes from that of ε, and is the result of our choice to consider
no rotations. Therefore, the present theory is inadequate to take into account the
effect of an externally applied density of torques, such as happens for magnetic
substances or very weakly connected materials (see Chapter 4), for instance. In
such a case, we have to enlarge the framework of our theory. This will be done in
the part devoted to Cosserat elasticity.
If ε contains only off-diagonal components, like for instance in a pure shear,
the elastic constant that comes into play is 2μ. This is why μ is also called the
shear modulus. In the case of an isotropic compression, ε is proportional to the
identity tensor. The elastic constant relating the pressure to the decrease of volume
is now (3λ + 2μ)/3 sometimes called «, the compressibility modulus - the factor 3
appearing in this expression is nothing but the space dimension, since tr(£ a ^) = 3.
Only two constants are needed to specify Hooke's law, when the solid is isotropic.
More generally, the stress field could have been introduced directly by a
linear relation of the form
&aß = Caßyß εΊβ, (3.7)
1. Elasticity 91

dw F
r 1 fr
; ! — m
w

? Li 1/
dl

Figure 3.2 An elastic solid under a uniaxial traction illustrates the two coefficients introduced
to characterize completely the elastic behaviour of a homogeneous and isotropic solid. The Young
modulus E is given by the ratio of the stress F/S and the relative extension dl/l. The Poisson
coefficient v is the ratio of the relative transverse contraction —dw/w and the relative extension.

where C is a fourth-order tensor, called the Hooke tensor. Relation (3.7) defines
the rheology of linear elasticity. Working with the latter tensor is rather tedious
and hides the simplicity of the theory. Therefore, we only keep in mind relation
(3.5) which is valid for isotropic solids.
Some other more complex, but still linear, relations can be introduced, in
non-local elasticity, where the stress field at point x depends on the strain tensor
at point y:
σαβ(χ) = J Οαβ7δ(χ, y) el6(y)dy. (3.8)

We will not consider this any further.


Inversion of equation (3.5) leads to

(1 + z/
) v
z ( \ (3.9)
εαβ — — ^ — & α β - -^θαβ\σΊ1),

where v and E are new constants called respectively, Poisson ratio, and Young
modulus. The relation between {E,v) and (λ,μ) can be found in the notation list
at the beginning of the book. The experimental interpretation of these constants
is shown in figure 3.2.
We can now rewrite the elastic energy in terms of the stress field as

1 (1 + v) 2 v λ2
(3.10)

As already mentioned only two out of the five elastic constants (E, v, A, μ, κ)
introduced up to now are independent. The positivity of the potential energy im-
poses that E, λ, μ and κ be positive. The Poisson ratio is different from the others.
92 S. Roux / Continuum and discrete description of elasticity

It is a dimensionless number. The positivity of £ implies the following bounds


— 1 < v < 1/2. In most materials, v is positive (around 0.3). The upper bound
(0.5) is reached for incompressible solids. Negative values are observed for com-
pacted sand.
Galilean invariance implies that the stress field should not do any work
during a rigid motion of the solid. This property can be used to derive the balance
equation satisfied by the stress field, always valid even if the rheological behaviour
considered is not linear. When the solid is subjected to an external force density,
/ , the balance equations read
δασαβ + fß = 0. (3.11)
Substitution of equations (3.1) and (3.5) into (3.11), allows to obtain a second-
order differential equation for the displacement field, known under the name of
Lame or Navier:
(A + μ)θαοβυβ + μοΊ0Ίυα + / β = 0, (3.12)
which needs to be complemented by boundary conditions to completely define a
well-posed problem.
In the above derivation, we have considered that at rest, U = 0, the solid
was stress free. Let us emphasize that this is not necessary. Most solids are in fact
not stress free at rest. The balance equations must be verified, but this does not
imply that σ is zero. However, since in general we do not know the initial state of
stress, σ0, we suppose that it is small compared to the one induced by the strains
studied.

To make contact with the simpler case of electrical conductivity, let us


rederive the basic equations for this case. We first postulate that the electric (dis-
sipative instead of potential) energy depends only on the first derivative of the
potential field, V(x), which is called the electric field, eQ = — daV. For an isotropic
solid, we are led to the following form of the energy:
S = |(fc(eeee)), (3.3')
2
where e = (eaea) is the only invariant, and k is a constant that we call the
conductivity. The conjugate to the electric field is the current, j a

Ja = "^ — kea. (o.o )

The flux of charges, Φ, through a surface dS is related to the current through a


relation similar to (3.6):
* = janadS, (3.6')
where n is a unit vector normal to the surface d 5 . The balance equation now is the
conservation of charge, and in the presence of a source term φ, it can be written
as
daja +¥> = <>. (3.11')
1. Elasticity 93

Finally, expressing the last equation in terms of the potential gives

kdadaV = φ. (3.12')

We hope to have convinced the reader that elasticity is not that different from
usual electrostatics.

3.1.3 Cosserat elasticity^


The displacement of a rigid solid in three dimensions can be characterized
by a translation, and a rotation; i.e. six degrees of freedom. In our description of
continuum elasticity, we introduced one displacement field, U(x), i.e. only three
degrees of freedom locally. The most general starting point consists in introducing
two fields: one for the displacements, U(x), and one for the rotations, φ(χ), of each
infinitesimal volume element. This will allow us to introduce "Cosserat elasticity",
the most general linear theory that fulfills Galilean invariance. We will see that the
model of the previous section is a particular case of "Cosserat elasticity", which
can be obtained by imposing an additional invariance. Two strains are introduced,
which should respect Galilean invariance:
• The strain: εαβ = daUß — €αβΊφΊ where e is the completely antisymmetric tensor.
The rotation field is subtracted to keep ε invariant under a rigid rotation.
• The torsion: καβ = οαψβ which is the gradient of the rotation field.
If we impose that φα = (l/2)e a ^ 7 ö^i7 7 , then the usual strain tensor ε is
recovered. More generally, this is one way to recover the theory of the previous
section as a particular case of the present framework.
Translations and rotations are not coupled in the expression of the potential
energy, as can be seen from the different symmetries of e and κ: Changing the parity
of space leaves ε invariant whereas n changes its sign. Since the energy must be
invariant under parity, there can be no cross terms in the energy. We introduce the
stress tensor, σ, and the moment tensor 1 , μ, respectively conjugated to ε and n.
The most general linear relations between stress and strains, for an isotropic solid,
are
ο-αβ = (μ + ο)εαβ + (μ - α)εβα + \εΊΊδαβ
μαβ = (7 + δ)καβ + (7 - δ)κβα + ηκΊΊδαβ
We have six elastic constants: μ and λ are the Lame coefficients; a would vanish if
the strain tensor ε were symmetric; the three last coefficients 7, δ and η are new.
The balance equations (analogous to eq. (3.11)) can be derived by studying
the equilibrium of an infinitesimal volume element. In the absence of external forces
and torques, the equations are daaaß = 0 for the stress and οαμαβ + €βΊασΊα = 0
for the moment. Combining these equations with the constitutive equations (3.13)
1
We use here the usual notations. However, the moment tensor should not be confused with
the second Lame coefficient which is also called μ.
94 S. Roux / Continuum and discrete description of elasticity

provides the differential equations satisfied by the displacement and rotation fields:

(μ + a)dßdßUa + (μ + λ - a)dadßUß + 2α€αβΊββφΊ = 0,


(7 + δ)Θβοβφα + {Ί + η- δ)Θαοβφβ + 2a€aßldßU7 - Ααφ = 0.

When φα = (1/2)€αβΊοβυΊ, we recover eq. (3.12) and its curl. Alternatively,


eq. (3.12) is also obtained if a = 0.
Among the six elastic constants, three have the dimension of a stress, λ,
μ and a, and the three others have the dimension of a stress times the square
of a length. By considering the ratio of these elastic constants, we can obtain
a length scale, i, which represents the scale at which local rotations take place.
Above this scale, rotations are no longer important. This property justifies the use
of the elasticity theory of the previous section. The restriction to only consider the
displacement field can be understood as a requirement for scale invariance. After
coarse-graining renormalisation, the additional terms of Cosserat elasticity vanish.
Which are the cases where Cosserat elasticity should be applied? We al-
ready mentioned the case of magnetic materials where a density of torques naturally
arises and therefore requires the use of a non-symmetric stress tensor. The same
applies to materials which have a microscopic structure and which are thus not
scale invariant. An example is given by granular materials where, at the scale of
the individual grains, local rotations are known to happen. Damaged materials
should also be treated in the same manner, at the scale of the microcracks. Finally
we also mention the case of vibrations where, if the frequency is large enough, the
length scale of the strain may become comparable to the intrinsic length I.

3.2 Other rheological behaviours


Up to now, we have only considered linear elasticity. Next we will consider
more complicated "rheologies" but we will maintain the notion of strain and stress
defined by their relation to the observable quantities of displacement and force
respectively.
3.2.1 Viscoelasticity[3]
Viscoelasticity is formally a straightforward extension of linear elasticity,
where now the stress state of a solid depends not only on the present strain, but
also on the past. We have previously mentioned the generalisation of Hooke's law to
a non-local relation in space (eq. (3.8)). The introduction of viscoelasticity requires
a similar extension but non-locally in time:
oo

<Taß(t) = J ΟαβΊδ{τ) εΊδ(ί - τ) dr. (3.15)


o
Elasticity is thus the limiting case where C is a Dirac distribution in time.
Purely viscous behaviour can be written as a linear relation between the stress
2. Other rheological behaviours 95

ε i σ A

Figure 3.3 Schematic responses of a viscoelastic solid. Note that in this example, the short-time
response is that of a viscous fluid, whereas the long-time behaviour is that of an elastic solid.
Some materials exhibit an inverse characteristic. Top: when a stress is imposed during a finite
interval of time (left), the strain can appear as shown on the right. Bottom: when the strain is
imposed (left), the resulting stress is shown on the right.

and the strain rate de/dt. It corresponds to C being the derivative of a Dirac
distribution in time. Figure 3.3 shows one typical response of a viscoelastic medium
under an imposed stress or strain.
Alternatively, one could also have introduced the compliance, L(T) (inverse
of the Hooke tensor C), by writing the strain as a function of the stress.
The long time behaviour of the tensor C(t) or L(t) is usually described by
an exponentially decreasing factor, with a characteristic time scale T, although it
may in reality consist of a large spectrum of intermediate relaxation times.
Since eq. (3.15) is a convolution, a Laplace transform in time yields formally
a relation similar to the one obtained previously for linear elasticity. Beside the
Laplace transform, it is also useful to consider the Fourier transform. Indeed, for
a harmonic force or displacement we can easily define a complex elastic constant,
whose real part corresponds to the elastic effect (stress and strain in phase), and
imaginary part to viscous effects (stress and strain with a difference in phase of
π/2, since the stress is proportional to the strain rate for a purely viscous fluid).
96 S. Roux / Continuum and discrete description of elasticity

B D

Figure 3.4 Schematic behaviour of a perfectly plastic solid. During an initial stage (OA) the
behaviour is elastic. When the stress reaches the plastic limit, σ ρ , A then a plastic flow takes
place, AB, and the stress remains constant. If the stress is decreased in B, then the behaviour is
again elastic (BC) but at zero stress a finite strain remains ε ρ .

3.2.2 Perfect plasticity^3'41


Plasticity is much more complex. It is characterized by an irreversible de-
formation that survives once a plastic strain has taken place, even when the forces
applied onto the system are reset to zero.
Let us first have a look at the example of figure 3.4 (top) where a simple
uniaxial tension is applied. For small stress, the behaviour is elastic (OA), and
no irreversible deformation occurs. When a threshold stress, the yield stress, σ ρ ,
is reached (in A) and if the strain increases, then a plastic flow will take place,
and the stress will remain at its threshold value. This simple case is called "per-
fect plasticity". Let us suppose that we reach point B and then decide to release
the stress. The stress and strain will follow the line BC, as if the behaviour was
elastic, with the same elastic modulus as in the first stage OA. However, once the
stress is zero, there exists a non-zero strain, ε ρ . This strain is called plastic and is
irreversible. If the strain increases again from C, we will follow the line CB, up to
the threshold stress where plastic flow will take place again (BD). The behaviour
is obviously non-linear since from B we can go toward C or D depending on the
sign of the strain.
How can we describe quantitatively this behaviour? In linear elasticity, we
have a relation between σ and ε, and thus once one variable is given, the other
can be simply computed. Here, for σ less than the yield stress, we can have any
desired strain. Therefore, we need to introduce another variable to characterize the
state of the material. We have seen that from any state (σ, ε) we have a linear
elastic behaviour when the stress is decreased to zero. We thus split the strain in
two parts: the "plastic" strain ερ that would remain if the stress was reset to zero,
2. Other rheological behaviours 97

and the "elastic" strain ee:


et = ε ρ + εβ. (3.16)
To avoid confusion, the strain ε is called the "total" strain et. We have seen that
at a given total strain many stress states could exist depending on the history of
the material. Therefore, we have to introduce an "incremental" law, which gives a
relation between time derivatives (denoted by a dot) of the strain and the stress.
We distinguish two cases:
1: If the σ is less than the yield stress, σ < σ ρ , or if σ = σ ρ and it < 0, the
incremental behaviour is elastic: έ ρ = 0 and ie = (1/Ε)σ. The total strain rate is
equal to the elastic strain rate as follows from eq. (3.16).
2: σ = σ ρ and it > 0. In this case, the stress will remain at its threshold, σ — 0,
and therefore the elastic strain will not vary, ee = 0. The total strain rate is equal
to the plastic strain rate.
Therefore, knowing the present state, σ, e e , ε ρ , and the total strain rate we
can compute the evolution of the stress and of each strain. This model is very simple
if only stresses in one direction are considered. It can be applied, for instance, to
a beam under simple traction.
Let us now turn to the three-dimensional case, and see how to generalize
the previous concepts. We first introduce the equivalent of the yield stress. Since
the stress is a tensor, we introduce a domain, X>, in the stress space inside which the
behaviour is elastic. For simplicity, we will characterize this domain by a function of
the stress, / ( σ ) . This function, called the "plastic potential", is such that / ( σ ) < 0
inside the domain X>, / ( σ ) = 0 on the boundary d P , and / ( σ ) > 0 outside. The
value of the plastic potential itself is not important for perfect plasticity, only its
sign is meaningful.
To properly define the plastic potential, we first need to make an exper-
imental observation: Under an isotropic pressure, most solids exhibit an elastic
behaviour for stresses orders of magnitude larger than the shear stress needed
to produce a plastic strain. Therefore, / ( σ ) should only depend on the part of
the stress, s(<r), for which the isotropic diagonal part has been removed so that
tr( s(cr)) = 0. More precisely, we introduce

8(σ) = σ - ftr(a)/ (3.17)

so that / does not depend on ίτ(σ). In isotropic media, / should be a function of


invariants of s: s n and sm defined in eq. (3.2) (si is zero by construction). In the
basis of the three eigenvalues of the stress tensor, Si, s2 and s 3 , the domain V is
a cylinder whose axis is parallel to the direction of an isostatic pressure (1,1,1).
Thus V can be defined completely by the intersection of this cylinder by a plane
perpendicular to the axis (1,1,1): Si -f S2 + «^ = 0.
The precise form of / is chosen according to experimental tests. However,
two forms of this function are rather popular because of their simplicity.
• The first one is known as von Mises potential. In the plane, «Si + s2 + ss = 0,
98 S. Roux / Continuum and discrete description of elasticity

the domain is a circle with radius, iZ, which is a material constant giving the yield
stress. The potential / is

/VM(<X) = y/ϊ^έη - R. (3.18)

It is easy to relate R to the yield stress introduced for, say, a uniaxial traction along
x. The only non zero component of the stress tensor is σχχ. Following eq. (3.18), we
have Si — 2/3σχχ and s2 = s$ = —1/3σχχ. Thus Sn = (2/3) σ\χ. On the boundary
of V f = 0 and therefore R = y ^ σ ρ .
• The second common choice for the plastic potential, due to Tresca, can be for-
mulated in a similar way. The elastic domain considered is again a cylinder, but
the section of this cylinder in the S\ + s2 + s 3 = 0 plane is a regular hexagon.
The potential depends on the largest eigenvalue of s that we write symbolically as
max(s):
/τ(σ) = max(i(j)) - R. (3.19)
In this case R = (2/3)σ ρ . Many other forms of the plastic potential are used,
depending on the material. In the following, we will only consider the von Mises
potential.
Next, we split the total strain into an elastic and a plastic strain as in
eq. (3.16). The spirit is the same as for the one-dimensional case: the*elastic strain
is related to the stress by Hooke's law. We rewrite eq. (3.7) in terms of rates:

σ = Cee
= C(et--ep). (3.20)

As before, two cases have to be distinguished:


1: If the stress is strictly inside the domain Z>, / ( σ ) < 0, the incremental behaviour
is purely elastic, and the plastic strain rate is zero ε ρ = 0. If the stress is on the
border of the elastic domain, dV, / ( σ ) = 0, we need to be more careful. Let us first
suppose that no plastic strain takes place: then we can compute σ from eq. (3.20).
Two cases can occur: either the vector σ points into the elastic domain - unloading
condition - and we face a situation similar to the first case (/ < 0), or
2: the vector σ is pointing out of the elastic domain - loading condition - and we
are in contradiction with the fact that σ should stay inside V. Therefore, plastic
flow will take place. There are many possible plastic flows that would be consistent
with all previous requirements. We introduce a new rule, called the "normality
law", to find a unique expression for ε ρ . This rule can be justified, but we simply
state it here. It expresses that the plastic strain rate will be parallel to the normal
of the surface dV. We introduce a variable β which measures the plastic strain rate
along the gradient of / in the stress space:

(3.21)
2. Other Theological behaviours 99

where ß has to be determined. Let us note that the loading condition is simply
ß > 0. In the uniaxial case, we required that the stress remains at its yield stress
value to obtain the plastic flow, for perfect plasticity. Here, we proceed along the
same lines, σ remains on the surface / = 0, so

(3.22)

which allows us to determine /?, using eqs. (3.20) and (3.21):

(3.23)

or
(3.23')

Once we know the total strain rate it, we can compute ß (eq. (3.23')) and thus the
plastic strain rate (eq. (3.21)). The elastic strain rate is then given by eq. (3.16).
Hooke's law (3.20) finally provides the stress rate.
Let us summarize the way to proceed: we suppose that the stress and the
strains are known at a given time step. Given the total strain rate we want to
compute the stress rate, and the elastic and plastic strain rates. If / ( σ ) < 0, the
incremental law is elastic (case 1). If / ( σ ) = 0, we compute β (eq. (3.23')). Ιί β is
negative, we are back to case 1, else we are in case 2. The procedure to follow in
the latter case has just been recalled above. Knowing the rates, we can integrate
one time step and repeat the procedure again. Therefore the problem is completely
determined.
3.2.3 Plasticity with work-hardening^3,4^
We have just considered the case of perfect plasticity where the plastic
potential is characteristic of the material and does not change with time. However,
in many plastic materials the elastic domain may change with the stress applied
to the system (see figure 3.5).
To take this into account, we need other information which defines how
the plastic potential, (the elasticity domain V) will change. This means in practice
other "degrees of freedom" in the description of the material. Once more, we will
consider the most simple cases, which can obviously be made more complex to
adjust more and more accurately to real experimental data. The two simplest
transformations of the elastic domain, P , in stress space, are 1) a translation and
2) a dilation. Keeping the generic form of von Mises plastic potential, we generalize
it to
(3.24)
We recall that in the basis spanned by eigenvalues of the stress tensor, the elastic
domain is a cylinder of circular cross-section. In the plane tr(cr) = 0, X is the
center of the circle, and R is its radius.
100 S. Roux / Continuum and discrete description of elasticity

Figure 3.5 Schematic behaviour of a plastic solid with work-hardening. In most plastic mate-
rials, the plastic limit - constant in the previous figure - increases with the plastic flow. This
phenomenon is known as work-hardening.

How do X, R and ε ρ vary as a function of time ? To answer this question,


we need to introduce a dissipation energy, £h, which will quantify how much energy
is dissipated by moving and/or growing the elastic domain in stress space. We in-
troduce the conjugate variables of X and R: dS^/dX and dS^/dR respectively. The
meaning of these quantities might seem rather abstract. To illustrate this concept,
we will use a metaphor: the conjugate variable to X is the "force" that opposes
the translation of V in the stress space. The conjugate to R is the "pressure" that
works against the swelling of V. £h gives the energy of the elastic domain as a
function of the "force" and "pressure". We emphasize that the words "force" and
"pressure" are used to illustrate the concepts introduced here, but have nothing to
do with a real force and a real pressure, in the same way that neither X nor R are
real distances, since we are in the stress space.
Since both X and R have the dimension of a stress, we know that their
conjugate variables will be dimensionally equivalent to a strain. Considering the
symmetries of X and R we can show that the conjugate variable of X is the plastic
strain ε ρ , and the one of R is the "cumulative plastic strain" p:

(3.25)

or p2 = tr(£p). Conjugate means that

X = ö4
(3.26)
R =
dp '
2. Other Theological behaviours 101

The time derivative of these equations gives X and R as linear functions of έ ρ and
P-
The following is similar to the case of perfect plasticity. If /(σ, X, R) < 0
or if /(σ, X, i?) = 0 with an unloading condition, then the behaviour is elastic: the
strain rate is purely elastic. All other parameters are constant: έ ρ = X = R = 0. If
/ ( σ , X , R ) = 0 with a loading condition, we write a generalized "normality law":
We introduce a parameter ß such that

P
6σ ÖX' (3.27)

The value ß is obtained by expressing that the stress remains on the surface / = 0,
i.e.
(3.28)

Using Hooke's law (eq. (3.20)), and the equations (3.26), we can express σ, X and
R as a function of the total strain rate, of έ ρ and of p and use these expressions in
eq. (3.28). Finally using the normality law, we obtain β as a function of the total
strain rate.
From this computation, we can extract all rates. First, we look at the sign
of β. If it is negative, we have an unloading condition, and thus the incremental
behaviour is elastic, all plastic parameters are constant: X, iZ, p, ε ρ . If β is positive,
then we obtain p and έρ by the normality law, ie by Hooke's law, X and R by the
dissipation equation (3.26).
So far, we did not specify the form of the dissipation. It can be fitted to
an experimental law, by computing the evolution in simple geometries. To our
knowledge, there does not exist a unique form of this potential that is of general
applicability.
3.2.4 Damage^
Some materials under stress are subject to a decrease of their elastic moduli
with or without showing signs of plastic deformations. This decrease of the elastic
moduli is due to the creation of microcracks in the material, and is generally called
"damage". Figure 3.6 shows a simple example of damaging behaviour with no
plasticity. We will describe this case following the same lines as for plasticity.
In figure 3.6, we schematize the behaviour of a damaging material under
uniaxial tension. We first increase the strain. The stress-strain relation follows the
bold line OAB. Let us note that the elastic modulus is continuously decreasing. If in
B we decrease the applied strain, the behaviour will be elastic, with no irreversible
strain - for zero strain we reach the origin 0 . When we increase again the strain,
up to point B, the elastic modulus is constant. After ε reaches ε<ι, in B, if the strain
continues to increase, then we will follow the bold line BD.
102 S. Roux / Continuum and discrete description of elasticity

Figure 3.6 Typical behaviour of a simple elastic solid with damage. Starting in O, when the
strain increases, the stress follows the bold line OAB. If the strain is released, the behaviour is
elastic (BO) with no plastic strain. The damage is the relative decrease of elastic modulus from
the initial state.

To describe the mechanical behaviour, we need to introduce an internal


variable which characterizes the irreversible evolution, i.e. the decrease of the elastic
modulus. Let us call E the initial Young modulus of the material, with no damage.
When some microcracks have been created in the material, the apparent Young
modulus is (1 — D)E. The parameter D, which goes from one to zero, is called the
"damage".
A very similar description applies to arbitrary loads. The classical descrip-
tion of damage relies on the following picture: the presence of microcracks screens
the stress in some parts of the material and increases the effective stress felt in some
other parts. As a first approximation, it is possible to capture part of this hetero-
geneity in the local stress distribution, by introducing an effective stress that would
produce the same strain with a non-damaged material. More precisely, at a given
stage of the damage characterized by a stress σ, there is a well-defined equivalent
stress σ which would give rise to the same strain, ε, in an intact (non-damaged)
material. Introducing Hooke's law for the damaged material:

a = Ce (3.29)

and for the non-damaged one:


& = Ce (3.30)

we can write

= (CC~l)a
(3.31)
3. Discretization: Physical approach 103

where we have introduced the fourth-order tensor £), named "damage". The hy-
pothesis of "equivalent stress" states that the properties of the damaged solid can
be modelled by that of the fraction of solid which supports the equivalent stress σ.
The damage is a fourth-order tensor which in simple cases can be approximated
by a scalar D times the identity tensor. The potential energy of the solid is thus
given by:
ε = (1/2)(1 - D)eCe. (3.32)
We also need to specify the boundary of the elastic domain, dV, inside
which the behaviour is elastic (i.e. curve OABD on figure 3.6). This can be done in
many ways. Analogously to what has been done for plasticity with work hardening
we could introduce a dissipation energy. For the case of uniaxial tension that we
discuss, it is easier to introduce the function, D = φ(ε). The damage, and therefore
the maximum strain for which the behaviour is elastic, can be described as a
function of the maximum strain £d that has been imposed on the solid in the past.
As for plasticity, we can describe the mechanical behaviour by an incremental law.
Here too, two cases have to be considered:
1: If we are inside the elastic domain ε < ε&, or if we are on the border ε = ε^ and
the strain rate is negative, then we are in the elastic regime:

σ = (1 - D)Ci. (3.33)

All other variables are constant: D — έά = 0.


2: If we are on the border of the elastic domain, e = ε&, and we continue to increase
the strain, έ > 0, then the damage will increase:

id = έ,
D = ψ'(ε)έ, (3.34)
σ = (l-D)Ci-DCe.
The two last equations can be simply integrated: D = ψ(ε^) and σ = (1 — ϋ)Οε.
We see the formal similarity between this damage theory and plasticity
with work hardening. Much effort has been spent to make more complex this last
rheological description, to take into account anisotropy and the full tensor character
of D, to couple the damage with plasticity and viscosity. The formalism of these
theories becomes quickly very heavy, but the spirit of the approach is very close
to the above presented naive model. Let us finally mention that a large number
of models exists to take into account the coupling between the damage and other
rheological behaviour, such as plasticity.

3.3 Discretization: Physical approach


We are now going to consider two different "philosophies" to implement the
general continuum rheological laws we have seen above, on a discrete system. The
first approach, which we call physical, consists in formulating an evolution law for
104 S. Roux / Continuum and discrete description of elasticity

a one-dimensional system, and then to construct a regular lattice, whose bonds are
governed by the one-dimensional laws derived. In other words, we require that the
elementary object considered (here a bond) corresponds to a "physical" concrete
object. The approach for a conductivity problem is similar to the one where we
build a network of resistors.
The second approach (called "formal") forgets about the local concreteness
of the elementary objects considered, and tries instead to describe mathematically
the continuum displacement field by the "best" approximation of this field onto a
simple basis of functions. This approach is discussed in the next section.
3.3.1 One-dimensional structures
The simplest example for a one-dimensional structure is provided by an
electrical resistor. We characterize its behaviour by the two potentials V\ and V2
at the ends of the resistor. Invariance with respect to the addition of a constant
to the potential imposes that all objective properties should depend only on the
difference Vi — V2. Linearity implies that the energy is proportional to (Vi — V2)2.
We thus obtain Ohm's law by taking the derivative of the energy with respect to
the potential. If we put these resistors on a lattice, the conservation of the current
at the nodes (Kirchhoff law), indicates that the potential at one node i, V^ is the
average of the potentials over the neighbouring nodes j ,
Vi = (l/z)YlVj, (3.35)
nn

swhere z is the coordination number of the lattice. One easily recognizes here the
discretization of the Laplace equation (eq. (3.12') without sources).
Let us try to work out a similar discretization for elasticity.^ For simplicity,
we restrict ourselves to a two-dimensional embedding space. Suppose first that we
only consider the coordinate of the displacement x and y at the two ends of one
bar, x being parallel to the axis of the bar, and y perpendicular to it (figure 3.7).
Invariance under translations imposes that only {χχ — x2) and (1/1 — 2/2) are relevant
for physical properties. If we imagine an infinitesimal rotation άθ at the first end of
the bar, the only change in the displacements will be for y2 which becomes y2 + id0,
where I is the length of the bar. Since the energy should not depend on d0, we
conclude that the only possible form for the energy is
E = \k(x2-Xl)2, (3.36)
This is exactly a linear spring. The independence of E on the y component means
that this spring is free to rotate at its ends. The only force supported by this bar
will be an axial one. This justifies the name of "central force" used to denote this
model. To express the same equation in a more intrinsic fashion (i.e. without the
need to have the x axis along the bar), we introduce the vector n along the axis of
the bar, and we call the two displacement vectors U\ and U2. The energy reads
Z = |fc{(i/x - U2) ■ n}\ (3.37)
3. Discretization: Physical approach 105

Figure 3.7 A one-dimensional elastic model. An elastic beam supports normal and shear forces
as well as torques. Each end is characterized in two dimensions by three coordinates: two for the
displacements, x and j/, and one for the orientation, Θ, of the tangent axis.

and thus the force into the bar is

F = k{(Ut - U2)n}n
= k{n x n)(l/i - U2) (3.38)

The relation between the force and the difference of displacements is given by the
elastic constant fc, times the projection tensor along the axis the bond.
If we want go one step further in the modelling of the one-dimensional
elastic structures, we can increase the number of degrees of freedom characterizing
the displacement of the bar. Thus we add angular variables, which give the local
direction of the bar with respect to the undeformed state. In two dimensions, only
one angle is necessary, but three are needed in three dimensions, and in general
d(d — l)/2 in a d dimensional space. We call θχ and θ2 the two angles, at the ends
of the bar. Since we want to build a linear theory, we require that x, y, and ίθ
are small compared to the length, £, of the bar. Invariance under translation and
rotation imposes that the energy depends only on (x\ — x2), {y\ — 2/2 + ^ 1 ) , {θ\—θ2).
In order to simplify the most general form of the energy, we consider bars which
are symmetric with respect to the x axis. In this case, since 8 should be invariant
under the change x —> —#, we obtain that the energetic contribution due to the
first term is decoupled from the others, thus

IE = A(x1-x2)2 + B{y1-y2 + te1)2


+ Cl(ßx - θ2)(υι -y2 + ίθχ) + Όί\θχ - θ2)\ (3.39)

If the bar is now symmetric under the exchange of subscript 1 and 2, we obtain an
additional relation:
C=-B (3.40)
therefore, three constants are needed to specify the form of the elastic energy.
Starting from very simple requirements of Galilean invariance and of lin-
earity, we have reconstructed the theory of elastic beams, commonly used in the
field of civil engineering since one century! To obtain the forces (conjugate to the
106 S. Roux / Continuum and discrete description of elasticity

displacements) and the torque (conjugate to the rotation) at node 1, we need to


differentiate eq. (3.39):

Fxl = A(xx - x2\


Bf
Fyl = B(yi - y2) +—(θ! + θ2), (3.41)
Bf
M1 = —fo-fc + ^ + D / ^ - f c ) .
The Galilean invariance automatically ensures that the bar is in static equilibrium.
Let index 1 and 2 refer to the two ends of the beam, then, one has for each beam:

Fxl + Fx2 = 0,
Fyi+Fy2 = 0, (3.42)
Ml + M2-Fylt = 0.
If one wants to describe a structure made out of beams where several beams can
join together at each point, the vectorial sum of the forces and of the moments
must be zero in equilibrium. This defines the behaviour of a network of beams.
Obviously, to continue the complete theory of beam elasticity we should go
to the limit of an infinitesimal length, and thus obtain local constitutive equations.
We finally connect this approach with the three-dimensional continuum elasticity
that we have introduced previously. We will not work along those lines hereafter.
We refer the interested reader to classic textbooks of strength of materials.
When the bar is embedded in a three-dimensional space, the equations
are more tedious but the concept is identical: the energy can be written into four
decoupled terms: the axial streching, the torsion, and the two flexions.
Formally, we can compare the approach we used in the continuum case and
in the discrete one-dimensional geometry. This may shed some light on Cosserat
theory of elasticity (section 3.1.3): in order to take rotations into account, we had
to enlarge the set of displacement variables. In turn, we had to incorporate elastic
constants which do not have the same dimension: A, B and D. Since in this case,
there is a well defined length scale, £, we used it to obtain comparable constants:
A, B and D£2. When the distance £ increases, we see that rotations and torques
(D£2 coefficient) are dominant in front of the shear (B coefficient). We recall that
in the bulk of a three-dimensional medium, when discussing the case of Cosserat
elasticity, we reached the opposite conclusion, namely, that torques and rotations
become irrelevant at a large scale.
If we now use these beams to construct a regular lattice, we have to consider
that the displacements and rotations at all the ends of beams which are connected
to the same node, are equal. As analogously, the Kirchhoff laws are nothing but
the balance condition at each node, in this case the sum of forces and the sum of
torques must vanish. Because of the nature of the local laws, we obtain a discretiza-
tion of a Cosserat elastic medium. This is in agreement with what we mentioned
3. Discretization: Physical approach 107

previously about this theory: the mesh size of the lattice naturally provides a local
microstructure. However, when the lattice size goes to infinity, this lattice effect
will vanish, unless some additional effect comes into play to reinforce it. This is
for instance the case for a random dilution such as for percolation (see chapter 4).
Most probably, this is also true for damage and fracture, since the microstructure
has a typical length scale which increases during the fracturing process.

3.3.2 Duality for elasticity in two dimensions


For electrical conductivity, in two dimensions, it is known that duality
transformations can be used. These duality transformations are valid in the con-
tinuum limit as well as in the discrete case. We can thus wonder whether such
a transformation can be applied to elasticity. The correspondence is, however, a
little more complex than in the conductivity case. We will begin by recalling the
result for resistor networks, and then we will describe the corresponding analogue
for elasticity. We will however skip all calculations, and just provide a qualitative
picture for it.
The first step consists in introducing all "admissible" fields of current in a
resistor network. A field of current is termed admissible if it satisfies the conserva-
tion of current at each node, and if it is compatible with the boundary conditions
when they are expressed in terms of currents. These admissible fields of currents
can be decomposed into loop currents - i.e. currents ji flowing around a loop i.
The actual current field is the one which minimizes the dissipated energy over the
set of all admissible currents. This property is similar to - in fact "dual" to -
the potential formulation where the potential field of the network is the one which
minimizes the total energy dissipated over all possible potential fields which satisfy
the imposed boundary conditions.
The duality transformation consists in exchanging the roles played by the
potentials and the loop currents. In this transformation, one also transforms the
original lattice into its dual, where all sites are transformed into loops and vice
versa, as shown in figure 3.8.
By this correspondence, the resistance of a bond becomes the conductance
of the dual bond. The voltage drop accross a bond (respectively the current flowing
through a bond) becomes the current flowing through (respectively voltage across)
the dual bond.
We can work out a similar transformation for elasticity, following the same
steps. We begin by constructing the admissible fields of stress. In the case of a beam
lattice, there are three independent stress fields per loop. These basic fields can be
imagined by introducing displacement jumps at the level of a loop, in the same way
as we can imagine a voltage jump (a generator) along a loop of resistors to produce
the loop current. These discontinuities are only a simple way to find the stress
distribution at equilibrium, but they do not correspond to any real discontinuity
(there is no generator along the loops of resistors in the electrical problem). The
three fields result from different discontinuities: one translation along the x axis, Ax,
108 S. Roux / Continuum and discrete description of elasticity

Figure 3.8 Starting from a lattice, shown by bold lines, we construct the dual lattice, shown by
dotted lines, by placing the dual sites inside the loops of the original lattice, and connecting these
sites by dual bonds which cross one and only one bond of the original lattice. In the particular
case of the square lattice, the dual lattice is also square.

one translation along the y axis, Ay, and finally a discontinuity in the orientation
of the beam, a "kink", B. The first two fields Ax and Ay consists in normal and
shear forces as well as torques, and one can be obtained from the other by a
rotation of π/2. The last field, B, consists in a constant torque along the loop. All
admissible stress fields can be expressed as a sum over loops of a linear combination
for each loop of these three fields. Here too the real stress distribution minimizes
the potential energy over all admissible stress fields.
The duality transformation, consists in identifying the three self-equili-
brated stress fields of one loop with the three components of the displacement of
the site dual to this loop. Some care should be taken in establishing the corre-
spondence. The only legitimate transformation is to interpret these displacements
as out of the plane. More precisely, Ax and Ay are angles out of the plane and
B is the displacement perpendicular to the plane as shown in figure 3.9. In this
transformation, one exchanges the elastic constants of an original bond and the
compliances of the dual bond.
It should be noted that these transformations are symmetric (the dual of the
dual is the original problem). Thus the out-of-plane displacements in the original
lattice map onto the in-plane displacements in the dual one. Remarkably, in-plane
and out-of-plane modes are completely decoupled but they are interchanged under
the duality transformation.
We have only considered the beam lattice case. In fact the duality trans-
formation also works for the central-force model on a triangular lattice. We map
again an in-plane deformation problem onto an out-of-plane one, and interchange
elastic constants and compliances.
Let us finally emphasize that these transformations in electrical or elastic
4. Discretization: Formal approach 109

Figure 3.9 The duality transformation for a beam lattice in two dimensions transforms an in-
plane deformation into an out-of-plane one, and vice versa. The three degrees of freedom per node
(two rotations and one translation) are shown here for the out-of-plane case. These displacements
correspond to admissible loop stresses in the original problem.

networks are also valid in the continuum limit. They are, however, specific for the
two-dimensional case. They have no equivalent in three dimensions.
3.3.3 Other rheological behaviours
We have seen the philosophy of the "physical" discretization approach at
work for elasticity. It can be done in a similar way for any other rheological be-
haviour. The first step is to describe the behaviour of a one-dimensional element.
Then, this description is used for all bonds in the lattice.
For viscoelasticity, we can directly reproduce what has been done in the
continuum formulation: i.e. introduce a relation between forces and displacements
at the ends of a beam which depends on time through a convolution with a retar-
dation kernel (see eq. (3.15)). The precise form of the kernel for the beam can be
computed from the three-dimensional relation by considering the stress distribu-
tion inside a single beam. Here too, a Laplace or Fourier transform gives a linear
relation between stress and strain, with no convolution.
Plasticity can be modelled along the same lines, although it is somewhat
more complex.

3.4 Discretization: Formal approach^


3.4.1 Finite element method for elasticity
We now turn to a very different approach for the discretization of the
continuum equations. We do not require any longer that the basic element used
has any physical meaning, such as the resistors or the beams introduced previously.
We rather choose here a more formal approach where the displacement field is
110 S. Roux / Continuum and discrete description of elasticity

approximated by a best fit within a subspace of trial functions. The method we will
describe below is called the "finite element method", and it is very popular because
of its simplicity (although very elaborate codes often use all sorts of sophisticated
optimizations) and its flexibility in handling very different problems.
We argued previously that the displacement field was the one that mini-
mized the total potential energy over all admissible fields. Let us suppose that we
do not explore the entire set of admissible functions, but only an affine subspace
of it. (An affine subspace is the set of functions which are the sum of a fixed func-
tion, and another one taken from a vector subspace. The fixed function is chosen
to match the imposed displacement on the boundary of the solid.) We can com-
pute the minimum of the energy in this subspace. In addition, since the energy
is quadratic in the displacements, finding the minimum in an affine space consists
in solving a linear set of equations. The use of the energy functional gives a well
defined "best" approximation. Moreover, to obtain a more precise estimate, we
only need to enlarge the subspace of trial functions over which the minimization is
performed. This is the essential ingredient of the method.
Now, we need to go a little more into the details of the finite element
method. Let us introduce the simplest form and then mention some extensions.
We recommend the interested reader to look for more details in textbooks for
applied mathematics or numerical computing. The first step is to divide space into
"elements" - triangles in two dimensions, or tetrahedra in three dimensions - in
such a way that no vertex lies on the edge or face of another element. The simplest
finite element method consists in considering the vector space of all functions which
are affine on each element (i.e. equal to a constant plus a linear function of the
space coordinate), and continuous over the whole domain. This choice is considered
for simplicity. A basis of this vector space is easily constructed. Let us call ψί(χ)
the basis function which takes the value 1 on vertex i and 0 on all other vertices.
Any displacement field U(x) can then be expressed as

tfW = E w W » (3·43)
i

where a, are the components of the trial function in the basis ψΐ(χ). For elastic-
ity where U(x) is a displacement field, a; are vectors, and each component of the
displacement field is decomposed into the same basis of functions. When a dis-
placement is imposed on a border, the values of the coefficients a, are fixed for
the corresponding nodes i. This last requirement selects an affine subspace of trial
functions.
We already see one advantage of this method. The division of space into
elements, also called "meshing", is very easy to build, and many codes provide
an automatic way of meshing a domain. The quality of the solution found at the
end of the computation depends, however, on the quality of the meshing, and
there exist sophisticated ways to reach a very efficient meshing or to adapt it
during the computation. These improvements are beyond the scope of the present
4. Discretization: Formal approach 111

introduction.
Next, one computes the energy of a trial function. This energy consists of
two parts: one is the elastic energy of the solid V:

£x = J a(U(x))e(U{x))dx (3.44)
v
and the other comes from the interaction with the external forces applied on the
system, which have a density f(x) in the bulk, and F(x) on the border dV

£2= I f(x)£(U(x))dx + / F(x)U(x)dx. (3.45)


v dv
The first part E\ is quadratic in the displacement field, since e(U(x)) is the sym-
metric part of the gradient of U(x), and a(U(x)) is related to e(U(x)) by the linear
relation eq. (3.5) which results from linear elasticity. The second part is linear
in U(x). With the decomposition of the displacement field in the basis of trial
functions we obtain the following expression for the energy:

\Σ,κνα*αά-ΈρΜ = 0' ( 3 · 46 )
where the matrix K is called the rigidity matrix :

Kij = \J σ(φί(χ))ε{Ψί(χ))άχ (3.47)


V

and the vector F represents the external forces:

Fi = J f{x)e{<pi(x))dx + j F{x)ipi(x)dx (3.48)


v dv
The energy functional eq. (3.46) is an extremum with respect each compo-
nent üi which is not fixed by the boundary conditions. We obtain the linear set of
equations
Kijüj = Ft (3.49)
that can be solved using standard numerical routines. Let us note that the rigidity
matrix is symmetric and positive, since the energy is positive. In addition, with our
previous choice of trial functions, it is a sparse matrix: K^ is zero whenever i and j
are not nearest neighbours nor equal, since either ψι or φ^ (or both) are zero on all
elements. When i and j are nearest neighbours or when i = j , the computation of
K^ can be easily done. It is a sum of integrals such as the one of eq. (3.47) where
ε(φ(χ)) and σ(φ(χ)) are constant over each triangle since φ(χ) is affine.
We have introduced the simplest elements. Many generalisations exist, ei-
ther for the shape of the elementary domains, or of the type of functions on these
112 S. Roux / Continuum and discrete description of elasticity

domains. Let us simply mention the use of second (or higher) order polynomials
as basis functions, which allows to have continuous test functions and first (or
higher) derivatives. It is also possible to add locally some special functions which
take into account some specific behaviour which cannot be accurately described
with the usual element. This is for instance the case at the tip of a crack where
it is well-known that the stress and strain fields diverge like 1/y/r where r is the
distance from the tip.
We also mention that the finite element method can be used to handle
problems of dynamical evolution or vibrations through minor adaptations of the
framework presented.
3.4.2 Extension to other rheological behaviours^
The previous approach can be adapted to deal with other rheological be-
haviours than elasticity. First, when the relation between stress and strain is non-
linear but still elastic, an incremental law can be used. At each step, the tangent
elastic modulii are computed according to the present state of stress or strain.
For plasticity, the situation is comparable to non-linear elasticity, although,
for elements which have reached the plastic limit, we need to know if the state of
stress will remain on the plastic limit (giving rise to plastic flow and eventually
work hardening) or on the contrary, if it will go inside the elastic domain (elastic
unloading). This is the non-linear aspect of plasticity. Provided that the modelling
is done incrementally, with a small enough time step, the problem can be handled
within the framework of finite element methods.
3.4.3 Comparison between the two discretizations
We should note that the stress field calculated with the optimum displace-
ment field through the finite element method, is not at equilibrium. More precisely,
it would be in equilibrium, if ever we had reached the exact solution, but in usual
circumstances, the latter does not belong to the space of trial functions. Look-
ing back at the discretized equations, we can see that the conjugate variables to
the degrees of freedom introduced satisfy the "balance" equations, like the one of
eq. (3.49). In some way, we have reconstructed an elasticity theory along the same
lines used in section 2, but for a finite number of degrees of freedom.
The two types of discretizations, the "formal" and the "physical" are not as
different as initially presented. The routes followed to obtain both sets of discrete
equations to describe a given continuum system are indeed very different. However,
both formulations respect Galilean invariance and linearity by construction. For
large system sizes (compared to the mesh size), and if the geometry of the medium
is itself homogeneous (e.g. not fractal) both systems are described by continuum
linear elasticity theory.
In the finite element method, the description of a real elastic solid will be
accurate only if the displacement field varies slowly over the size of the elements
used. In particular, if we want to introduce a crack in a medium, the displacement
References 113

field around the crack will be accurate only if we use elements much smaller that
the size of the crack itself. This is a serious drawback for the modelling of a dense
array of microcracks in a medium.
Similar drawbacks, although less dramatic, will appear if we want to model
a heterogeneous solid. The size of the heterogeneities should be large compared
to the typical size of the mesh. Fluctuations of elastic modulii from element to
element should be small.
On the other hand, the models obtained from the "physical" approach
are realistic, in the sense that they accurately describe structures that can be
built. Therefore, the general laws we can extract from a variety of properties of
heterogeneous systems are meaningful. In addition, one recovers naturally in the
beam models the full "Cosserat" elasticity, which may be important for the cases
where the microstructure of the medium evolves such as for rupture. This apprach
has also many drawbacks: No elastic solid can be modelled by a lattice of beams
or central-force bars. In particular, an incompressible solid cannot be described in
this way.
We therefore suggest the following use for these two approaches:
• In order to study a general physical behaviour involving a complex microstruc-
ture, in particular with heterogeneities, the "physical" approach is more suitable.
Within this framework, one can study the statistical behaviour, and see its evolu-
tion with the system size. The laws observed, in the thermodynamic limit, i.e. for
an infinite system should correspond to realistic laws and this, a posteriori justifies
the use of this approach in many of the cases studied in the following chapters.
• If the aim is to describe a given homogeneous solid, then the "formal" approach
is certainly more suitable. However, it should be used in such a way that the mesh
size is much smaller than any typical size involved in the geometry of the sample.
This last requirement makes difficult to handle this approach when dealing with
systems containing a large number of heterogeneities.

References
1. A.E.H. Love, A Treatise on the Mathematical Theory of Elasticity, (Dover,
New York, 1944);
L.D. Landau and E.M. Lifschitz, Elasticity, (Pergamon Press, New York, 1960);
W. Prager, Introduction to Mechanics of Continua, (Ginn, New York, 1961);
L.I. Sedov, Introduction to the Mechanics of a Continuum Medium, (Addison-
Wesley, Reading, 1965)
2. W. Nowacki, Theory of Micropolar Elasticity, (Springer-Verlag, Udine, 1972)
3. F.A. McClintock and S.A. Argon, Mechanical Behavior of Solids, (Addison-
Wesley, Reading, 1966);
F.R. Eirich, Rheology, (Academic Press, New York, 1956);
J. Lemaitre and J.L. Chaboche, Mecanique des Materiaux Solides, (Dunod,
Paris, 1985)
114 S. Roux / Continuum and discrete description of elasticity

4. E.H. Lee and P.S. Symonds, Plasticity, (Pergamon Press, New York, 1960)
5. B. Wilshire and D.R.J. Owen, Engineering Approaches of Engineering Design,
(Pineridge Press, New York, 1983);
Z.P. Bazant Advance Topics in Inelasticity and Failure of Concrete, (Cement
Och. Betoninstituted, Stockholm, 1979)
6. S. Timoshenko, Strength of Materials, (D. Van Nostrand Co., Princeton, 1955);
J.P. Den Hartog, Strength of Materials, (Dover, New York, 1949)
7. P. Tong and J. Rossetos, Finite Element Method, (MIT Press, Cambridge,
1977);
P.G. Ciarlet, The unite element method for elliptic problems, (North-Holland,
Amsterdam, 1978);
D.H. Norrie and G. de Vries, The unite element method, (Academic Press,
New York, 1973)
8. K. Washizu, Variational methods in elasticity and plasticity, (Pergamon Press,
New York, 1975)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
©Elsevier Science Publishers B.V. (North-Holland), 1990 115

4. Disorder

Alex Hansen*

4.1 Introduction
Rarely are materials of practical interest perfect. Not only is it difficult to
make such perfect materials, but often the presence of disorder is desirable. It has
often a profound influence on the material properties. One good example of this is
concrete. As was described in section 2.3, the strength of this material - and thus its
usefulness as the prime building material - is due to its inherent disorder. Another
example is the annealed copper pipes used by plumbers. These pipes are very soft,
easy to bend into the shape to fit beneath the kitchen sink. The bending process,
however, produces an interlocked tangle of dislocations in the crystal structure of
the copper, making the pipes rigid and durable.
Disorder is a broad concept. In the first of the two examples above the
disorder comes from the material being a composite on a mesoscopic scale - i.e. a
scale much larger than the atomic one, the microscopic scale, but still, we assume,
much smaller than the size of the sample itself, the macroscopic scale. In the
second example the disorder is in the crystal structure of the copper, i.e. on the
atomic scale. The methods used to theoretically analyze the effect of disorder on
the material properties, are of course dependent on the kind of disorder that is
involved. However, a few concepts and ideas are common. One of these is that of
homogenization.
A disordered material is by definition heterogeneous: Its local material
properties vary spatially. Suppose now that it is possible to define a representative
volume element, which is small compared to the dimensions of the sample, but
much larger than the length scale of the disorder (which is the correlation length
to be defined below). If the material is statistically homogeneous, which means
that the local material properties are constant when averaged over a representa-
tive volume element, then it is possible to replace the real disordered material by a
t Fysisk Institutt, Universitetet i Oslo, Postboks 1048 Blindem, N-0316 Oslo 3, Norway.
116 A. Hansen / Disorder

Figure 4.1 We calculate the conductance of this network via the mean field approximation. The
conductance σ varies from bond to bond.

homogeneous one, where the local material properties are the averages over the rep-
resentative volume elements in the original material. This replacement constitutes
the homogenization process.
When a homogenization has been carried out, the homogeneous model
material is used as a starting point to calculate whatever macroscopic properties
one is interested in. One should, however, notice that different properties have a
vastly different sensitivity to the disorder. For example, transport properties such
as electrical conductance are not sensitive in comparison to material strength, which
is governed by the catastrophic growth and merging of unstable microcracks.
The program of homogenization as stated above seems simple. It is, how-
ever, not. The difficulty lies in finding the local averages of the material properties.
A tremendous effort has been devoted to various ingenious schemes to calculate
these averages. A nice review of some of these methods is that of HashinJ1^
Let us go through a simple example of a homogenization procedure, to fully
demonstrate the ideas involved, since this technique is so central to most treatments
of disordered materials. The example we have in mind is Kirkpatrick's calculation
of the conductance, Σ, of a network of conducting bonds with a distribution of
conductances.^ The result we will obtain will also be useful for a comparison with
the results from percolation theory, which we present later in this chapter. This
particular homogenization procedure is an example of what physicists refer to as
a mean-field approximation.
The network is a quadratic section of a square lattice, and is shown in
fig. 4.1. Its upper and lower boundaries are connected to bus bars kept at a voltage
difference V. The lattice is tilted at a π/4 angle compared to the boundaries of the
lattice. With this arrangement all bonds carry the same current if the conductances
are all the same, i.e. in the case of a homogeneous lattice. The homogeneous network
1. Introduction 117

k if / ^
-^—^
A i/2
' 11

Figure 4.2 The setup used to calculate the conductance between nodes A and B.

we have constructed consists of bonds with a conductance am. The voltage across
each bond in this homogeneous lattice is then vm = V/L, where L is the linear size
of the network. The current flowing through each bond is zm. The relation between
this current and the voltage difference across each bond, v m , is found by calculating
the conductance across a single bond in the homogeneous network: Suppose the
network is very large and grounded at the edges of the network (thus setting the
voltage V to zero for the time being). We connect a wire between ground and
one of the two nodes attached to some bond in the network. Let us call this node
A. A potential difference between ground and this node is set up, and a current
i flows into the network through this node. Symmetry demands that a current
z/4 flows through each of the four bonds attached to node A if L is very large
and A is far from the edges of the network. Let us now move the wire to another
node connected to the chosen bond, which we call B. We also reverse the potential
difference between the node and ground in such a way that a current i flows from
the network through the attached wire. This arrangement is shown in fig. 4.2. Of
course, the current through each of the four bonds connected to this node is also
z/4 when L is large and B is far away from the edges of the network.
Since the equations governing the currents flowing in the network - the
Kirchhoff equations - are linear, we may superpose these two situations. Thus
we find that the current flowing through the bond connecting nodes A and B is
i/4 + z/4 = i/2. The conductance of the bond is a m , so that the voltage difference
across it is i/2am. The conductance between the two nodes is then the current
flowing into the network, i divided by the voltage difference between the nodes,
i/2am. This gives
σ ΑΒ = 2a m . (4.1)
Let us now go back to the original, disordered network. We focus our at-
tention on a bond with conductance σ. With the external voltage difference V
imposed across the bus bars, the bond sustains a voltage difference vm + v, where
v is the deviation from the mean value vm.
118 A. Hansen / Disorder

The mean field approximation consists in replacing the rest of the net-
work by the homogeneous network, but keeping this particular bond unchanged -
an approximation amounting to ignoring the correlations between the fluctuating
voltage v and the fluctuating conductances. The conductance σΑΒ between the two
nodes connected to this bond, let us once again call them A and B, is easily found:
The bond with conductance σ, is connected in parallel to the rest of the network.
By using the result for σ'ΑΒ in eq.(4.1), we find

ΟΆΒ = °'AB ~ σ™ + σ = am + σ. (4.2)


Thus, the current-voltage relation across this bond is

{vm + v)(am + σ) = z, (4.3)

or solving for the fluctuating voltage

v = i/(^ m + σ) - vm. (4.4)

The average of the fluctuating part of the voltage over the network must be zero:
(v) = 0 = (i/(am + σ)) — vm. In the average on the right hand side of this equation,
we ignore once again the correlations between fluctuating conductances and the
resulting voltages, or in this case currents:

(i/(am + σ)) = <i)(l/(* m + σ)) = 2amvm(l/{am + σ)). (4.5)

Thus, we end up with the equation

0 = < ( * m - * ) / ( * m + a)). (4.6)

Let us now suppose that the distribution of conductances is given by the probability
density ρ(σ) = ρδ(σ — σ0) + q6(a), where δ(χ) is the Dirac delta function. The
meaning of this probability density is that a proportion p of bonds are present in
the network, while a proportion q = 1 — p are absent (i.e. having zero conductance).
Equation (4.6) then becomes
1
0=1 [p(a){am - a)/{am + σ)] άσ = p{am - a0)/(am + σ0) + q. (4.7)
o
Finally, solving this equation for am gives

am = a0(l-2q) = 2a0(p-l/2). (4.8)

We have now found the average value of the local conductance, am, required
for the homogenization process. It is now completely straightforward to find the
conductance between the bus bars, since we are now dealing with a homogeneous
lattice: It is simply
Σ = am = 2σ0(ρ - 1/2). (4.9)
2. Percolation model 119

This finishes our example. Notice, however, that eq. (4.9) predicts that the conduc-
tance of the network is zero for p = 1/2, and changes sign for p less than this value.
This change in sign is obviously unphysical, signalling that the approximations are
no longer valid. Is this a breakdown of this particular scheme, or a breakdown
of the statistical homogeneity assumption underlying the homogenization process?
The latter is actually the case here.
The reason for this breakdown may be understood intuitively as follows:
When a proportion p bonds are present and a proportion q are absent, one may
define clusters of connected bonds with zero conductance. Connectedness in this
case means the following: Pick two bonds with zero conductance. If it is possible
to draw a path along the network between these two bonds passing only through
zero-conductance - or absent - bonds, only then the two bonds are connected. A
cluster is a set of all mutually connected bonds. Now, it is possible to regard the
network as a composite, where, when the network is conducting, the conducting
bonds form a matrix and the clusters of non-conducting bonds form the particles.
The fundamental assumption necessary for the homogenization procedure is that
representative volume elements may be formed. However, when p decreases, the size
distribution of the clusters of non-conducting bonds becomes broader and broader.
When p & 1/2 the distribution is so broad that no representative volume elements
may be constructed.
This broadness of the cluster size distribution, n s , which is the number of
clusters of size s per site, is the starting point of a completely different approach to
handle disorder compared to the homogenization approach we have been discussing
so far. This is the percolation approach, which we will discuss in the next section. As
opposed to the homogenization approach, it is valid when the disorder is so strong
that the length scale above which the material may be regarded as homogeneous
is large compared to the microscale of the system.

4.2 Percolation m o d e l
In the following we will discuss the percolation approach only in terms of
lattice models, and refer the reader to the literature on the subject of continuum
percolation}3^ An excellent book on the subject of this section is given in ref. [5].
Percolation theory started 1957 with the publication of a paper by Broad-
bent and HammersleyJ6] They showed rigorously that a network, such as the one
discussed in the previous section, will not conduct if the concentration p of con-
ducting bonds is smaller than some non-zero pc. It is obvious that the network will
not conduct as p —» 0. However, that its conductance drops to zero before p is zero,
is less evident. The consequence of this is that the conductance, Σ1, of the network
as a function of p cannot be an analytic function for all p; loosely put, if an analytic
function is constant somewhere, it must be constant everywhere. Thus, the con-
ductance Σ is singular for some p = pc. This pc is called the percolation threshold,
and is recognized as the value of p for which the conductance first becomes zero as
120 A. Hansen / Disorder

p is lowered.
If we go back to the mean field result of eq. (4.9), we note that this ex-
pression predicts a vanishing Σ for p = 1/2, and a non-physical behaviour for p
less than this value. Since, the conductance cannot increase as p is lowered, the
unphysical expression for Σ should be substituted with Σ = 0. Thus, we see that
the mean field calculation gives, at least qualitatively, a correct picture of what is
happening. In particular, the singularity in Σ is predicted by the breakdown of the
calculation.
So far we have limited our discussion to the square lattice. The existence of
a percolation threshold is, however, quite general. Let us now define more generally
the percolation model: We have a d-dimensional infinitely large regular lattice of
some topology. It may for example be a three-dimensional cubic lattice, or a two-
dimensional triangular lattice. Let us now distinguish between bond percolation
and site percolation.
We define bond percolation first. With probability p a bond is "present"
(according to some definition such as having a non-zero conductance) or is absent
with probability q = 1 — p. We define a cluster of present bonds as the set of all
present bonds connected to each other. Thus, picking any two bonds belonging
to the same cluster, it is possible to find a path between them only passing over
present bonds. Now, if there exists a cluster containing an infinite number of bonds,
the network percolates. Otherwise it does not. The connection with the above
discussion of the conductance of the square lattice is as follows: For the network in
that example to conduct, there must be a cluster of conducting bonds large enough
to span the network from bus bar to bus bar. Above the percolation threshold p c ,
such a spanning cluster will exist, also in the limit of L —> oo. For p less than pc
such a spanning cluster does not exist in the limit of L —> oo.
In site percolation, the nodes - or sites - rather than the bonds are the
elementary blocks of the lattice. A site is either present with probability p or absent
with probability q = 1 — p. To define clusters in this case, we proceed as in the
bond percolation case: If we can find a path between two sites that does not pass
through any absent sites, these two sites are said to be connected. A cluster is a set
of all sites which are all mutually connected to all other sites within this set. Also
for the site percolation problem there exists a well defined percolation threshold,
which separates a phase where no infinite cluster exists from a phase where one
does exist.
In table 4.1 we list the percolation thresholds for bond and site percolation
for some different lattice types. Particularly, one should note the value pc = 1/2
for the two-dimensional square lattice. This is an exact result (which we will derive
in section 4.3). Notice that this value is the same as the one found in the mean
field theory. This is accidental. The general result for pc from mean field theory
for bond percolation on any lattice type is pc = 2/z, where z is the coordination
number of the lattice. From table 4.1, we see that the mean field values and the
correct values for the thresholds approach each other as the coordination number
2. Percolation model 121

of the lattices increase. This is an indication that the mean field theory becomes
better and better with increasing coordination number.

site bond mean field


square 0.59275 1/2 1/2
triangular 1/2 2sin(7r/18) 1/3
honeycomb 0.6962 l-2sin(jr/18) 2/3
cubic 0.3117 0.2492 1/3
bcc 0.245 0.1785
fee 0.198 0.119
diamond 0.428 0.388
Table 4.1 The site and bond percolation thresholds for some different lattice types. We also
include the mean field values of the bond percolation thresholds for some lattices.

At about the same time as Broadbent and Hammersley introduced per-


colation theory, the statistical physics community was beginning to focus on a
particular kind of phase transitions known as second order phase transitions or
critical points. In fig. 4.3 we show a sketch of the phase diagram of a real gas.
P is pressure, V is volume, and T is temperature. The point marked "C" is the
critical point. It has coordinates (PC,VC,TC). If the temperature is above Tc there
is no pressure for which the gas liquidifies. Below Tc there is. In the vicinity of
the critical point the gas develops what is known as critical opalescence: it looks
milky. The reason for this is that gas contains liquid droplets with a very broad size
distribution, and those droplets with sizes close to the wavelength of the incoming
light strongly reflect and refract it. The existence of critical points is actually fairly
common in the phase diagrams of various physical systems. Throughout the six-
ties, a general scaling theory was developed for the behaviour of physical systems
near critical points. This effort culminated in the renormalization group theory of
Wilson in the early seventies. Through this theory it became possible to analyze,
and quantitatively correctly predict the behaviour of the systems near the critical
points. An entry point to the vast literature on this subject is provided by the
series of books edited by Domb and Green. ^
It happens that the percolation threshold pc is a critical point in the same
sense as described above. Thus, all the theoretical machinery developed to tackle
second-order phase transitions is directly applicable to the percolation model. Thus,
as we now go on to present the scaling theory of percolation, the reader should
have in mind that these results are quite general and applicable to a large variety
of different physical systems. We will in the following assume a bond percolation
problem.
We are again assuming an infinitely large regular lattice. The basic quantity
of interest in the percolation problem - at least as far as geometrical properties are
concerned - is the cluster size distribution n s , giving the distribution of clusters
122 A. Hansen / Disorder

T>TC
T = TC
T<Tr

Figure 4.3 Sketch of the phase diagram for a real gas. The point marked UC" is the critical
point.

of size s per bond - which is defined as the probability that a bond belongs to a
cluster of size s. Normalization demands that

q + ]T sn8 + Poo = 1, (4.10)


finite 8

where P ^ is the probability that a bond belongs to the infinite cluster. Above p c ,
Poo is non-zero, while for p less than p c , it is identically zero. This is illustrated in
fig. 4.4. Such a quantity is generally referred to as an order parameter. The density
of the infinite cluster is also a non-analytic function of p, just as the conductance
is. It has a singularity at pc. Through eq. (4.10) we see that the sum Zenite * 5 n *
must have a similar singularity in p at pc as the density of the infinite cluster, Ρ ^ .
This singularity can be described by a non-integer critical exponent β, given by

Poo = Ai(p — pcY + less singular terms, (4.11)

near the critical point. Other moments of the cluster size distribution also show
singular behaviour near pc. For example,
a
Σ nM = A0{p-pc)2 + ···, (4.12)
finite s

and
7
Σ s2n8 = A2(p-pc) + (4.13)
finite 8

where a and 7 are critical exponents.


The critical exponents have a very remarkable property. Their values are
only dependent on the spatial dimensionality of the lattice, and not on the lattice
2. Percolation model 123

0 Pc 1

Figure 4.4 The density of the infinite cluster, P^, as a function of the probability for a bond to
be present, p. P ^ is an example of an order parameter, and p is an example of a control parameter.

type or whether we are dealing with bond or site percolation. This property is
known as universality. In ordinary critical phenomena the same amazing property
is found: there are only a few different universality classes. All systems belonging
to the same universality class share the same critical exponents. This leads in prac-
tice to, for example, such diverse systems as magnets and liquids having identical
critical exponents. In table 4.2 we list some critical exponents for percolation.
Another quantity of great interest is the connectedness correlation function
G(r), which is the probability that two bonds a distance r apart are connected.
Also this quantity has a singularity at pc, and its behaviour near this point is
e-r/t(p)
G(r) ~ -μζ^, (4-14)

where η is yet another critical exponent. However, notice the length scale ξ — ζ(ρ)
appearing in the exponent. This is the correlation length, which diverges with a
critical exponent v near pc:
£~ΙΡ-ΛΓ. (4.15)
The correlation length is an important quantity. The reason is the following: At
pc, the correlation length is infinite, and the exponential in eq. (4.14) is one. This
means that there is no length scale in the correlation function G(r). This is gen-
eral: At pc there is no length scale in the problem (except the lattice constant).
Away from pc (but not too far) the only additional length scale appearing in the
problem is the correlation length ξ. This enters through the exponential appearing
in eq. (4.14). Thus, if we study the network on a length scale smaller than the
correlation length, it will look as if we are at pc. At length scales larger than pc,
the network appears homogeneous. Let us now present a typical scaling argument
which makes it possible to relate the various exponents to each other. We go back
124 A. Hansen / Disorder

d=2 d=3
V 4/3 0.88
ß 5/36 0.44
Ί 43/18 1.76
τ 187/91 2.20
σ 36/91 0.45
t 1.300 2.02
s 1.300 0.75
Τ 3.96 3.8
S 1.30
f 4.0
9 1.30
Table 4.2 Some critical exponents for percolation for two- and three-dimensional lattices.

to the cluster size distribution, n s , and ask what is its behaviour at pc. It turns out
that it is a power law in s,
n8 ~ s~T. (4.16)
Likewise, the radius of gyration averaged over the cluster of size s, i2 s , is also
singular,
R8 ~ 8σν. (4.17)
We now ask what does n8 look like away from pc. If s is so small that R8 is smaller
than the correlation length £, n8 still behaves as in eq. (4.16) - there cannot be
any dependence on p. However, for s such that R8 is larger than £, the cluster
size distribution falls off very rapidly.1 We may use this information to write the
behaviour of n8 as a function of two variables, s and R8/£:

ns = s-Tf(Rs/aP)) = s~Tf((P ~ P c K ) , (4.18)

where / is a scaling function, f approaches a constant as its argument goes to zero.


For large arguments it falls off very rapidly. Let us now rewrite this equation,

n8 = s-T/((p-pcK)
= (P - Ρο)τ/σ((ρ - PcVrT/<Tf((p - PcK)
= (ρ-ΡοΥ/σ9((ρ-Ρο)8σ\ (4.19)
1
The behaviour of ns for clusters with very large s so that their radii are much larger than
the correlation length, is
n8 ~ s~e exp(-cs^).
For p > Pc, one has that Θ is 1 in two dimensions, and 3/2 in three dimensions. The other
exponent, ζ, is 1. This is known as the lattice animal limit Above p c , Θ — 5/4 in two dimensions
and —1/9 in three, ζ is also different: ζ = 1 — 1/d where d is the spatial dimension.
2. Percolation model 125

where g(x) = x T^f(x). We calculate the nth moment of the cluster size distribu-
tion:

£ snn, = (p-pcy"Jdssng((p-Pc)s°)
finite *

= {ρ-Ρο)τ/σ~{η+1)/σ Jdxxng(x). (4.20)

Comparing this equation with eqs. (4.11), (4.12), and (4.13) we find that,

2 - α = β + 1/σ = - 7 + 2/σ = (τ - 1)/σ. (4.21)

Thus, all the critical exponents describing the moments of the cluster size distribu-
tion may be related to the two exponents r and σ. But what about the exponents
appearing in the correlation function G(r), eq. (4.14)? They may also be related to
r and σ, but the arguments leading to these relations are different than the scaling
arguments presented above.
The first of the two arguments goes as follows: Set p slightly different from
pc. The second moment of n s , eq. (4.12), is the same quantity as the integral of
the correlation function over the entire (infinite) lattice, because both measure the
average number of bonds connected to any given one. The correlation function falls
off rapidly for sizes larger than the correlation length £. Thus, we may approximate
this integral as follows:

j ddr
J rd-2+V
~ ^""-(P-Pc)^2-"»". (4.22)

This result is then compared to eq. (4.12), and we find

(2 - 17)1/ = 7 . (4.23)

The second argument is known as a hyperscaling argument: suppose we


have a correlation length £. Then a typical cluster has a radius of the order of £.
This cluster contains £dPoo(p) bonds in d dimensions. Thus,

ξ"(ρ - Pcf ~ (P - Pc)0-"" ~ n ; ^ I P o o ~ ((p - Ρ ο ) τ / Τ 1 / τ , (4-24)

where we have used eq. (4.18). This leads to

β + 1/σ = du, (4.25)


126 A. Hansen / Disorder

Figure 4.5 The Cantor set. It is constructed by cutting away the central third of a line segment
of length one. The central third of each of the two segments that are left are then removed, and
so on. The set one is left with after this process has been carried out an infinite number of times
is the Cantor set. If we take the left third of the set and enlarge it by a factor three, the original
set is recovered.

which is the hyperscaling relation. 2 Thus, we are back to having only two critical
exponents which must be determined by methods outside the scaling theory for
critical phenomena. No matter the number of exponents we define to describe the
connectedness properties of the percolation model, the number of independent ones
is two.
At this point it is natural to make a connection with the concept of fractals
- which has proven to be immensely valuable in quantifying geometrical shapes
which earlier were dismissed as "too complicated". The reader will find the book
by Feder a nice starting point for a serious study of this subject.^
The central feature of fractals is their self-similarity - that is to say, fo-
cusing on a piece of some fractal, enlarging it to the size of the original fractal
makes an identical copy of the original. In fig. 4.5 we show a simple fractal known
as a Cantor set. If a part of it is cut out and enlarged by some integer power of
three, the original set is recovered. However, this scale invariance is also found
for, say, line segments. Thus, there is more to fractals than this property alone.
What distinguishes fractals from these more mundane objects, is that they have a
non-integer fractal dimension. Let us now define the similarity dimension, which is
closely related to the fractal dimension but simpler to grasp and to use in practice. 3
2
The hyperscaling argument works only in less than or equal to six dimensions! The other
scaling arguments are always OK. If we had been more careful in stating the assumptions behind
the scaling and hyperscaling arguments, we would have found that it is only in the last one that
we have explicitly used that there is only one length scale, £.
3
The fractal - or HausdorfF-Besicovitch - dimension is defined in the following way: we want
to measure the "mass", M of our object. We do this by covering it with balls of diameter less
than or equal to e. The mass of the balls is proportional to their volumes, which in turn are
proportional to ed where d is a dimension. The mass of the entire object is the sum of the masses
of the balls. Among all the possible coverings, one first chooses the ones that have the smallest
number N(d) of balls, and among those coverings one chooses the one with the smallest mass.
The HausdorfF-Besicovitch dimension, D, is then defined by the following expression:

if d > D;
M = N(d)ed -+
{ if d < D,
as e —» 0.
2. Percolation model 127

For our purposes here we do not need to distinguish between these two dimensions.
If we cut a line segment of length L into segments of length /, we get N = L/l new
segments. Cutting a square with sides having length L into squares with sides hav-
ing length /, gives N = (L/l)2 copies. In general, cutting a d-dimensional cube with
sides length L into cubes with sides of length /, produces N = (L/l)d smaller cubes.
The Cantor set of unit length in fig. 4.5 produces N = 2 copies when cut into pieces
of length / = 1/3. We relate N and / as we did for the cube: N = 2 = (1//) D = 3 D ,
where D = log 2/ log 3 « 0.6309. D is the similarity dimension for the Cantor set. It
is non-integer. Thus, the Cantor set is a fractal, with fractal dimension log 2/log 3.
The Cantor set is an example of a deterministic fractal. Naturally appearing
fractals are typically statistical fractals, i.e. their self-similarity is of a statistical
nature. The meaning of this is the following: We have an ensemble containing all
possible realizations of some statistically fractal object. If we pick out one typical
realization, cut out a piece of it and rescale the size of this piece to the size of the
original object, we find that after the rescaling, the new object is identical to some
typical member of the original ensemble.
The definition of the similarity dimension for a statistical fractal may be
given as
M
<L> (L\D ,i,fil
(4 26)
MtfpUV ' ·
where M(L) is the mass of the fractal object on the length scale L averaged over
the statistical ensemble of realizations, that do not have zero mass. The ratio
M(L)/M(l) reduces to TV, the number of copies in the case of deterministic fractals.
The fractal dimension of the infinite cluster in the percolation problem is
easy to relate to the critical exponents we have defined. The "mass" M(£) of the
infinite cluster within a volume £d, i.e. the number of bonds belonging to it, is
given by eq. (4.24), and by using eq. (4.15), we find

M ( 0 ~ £dPoo ~ £,d{p - Pcf ~ ξ"-β/ν. (4-27)

Cutting the volume of size £ into pieces of size / equal to the lattice spacing,
and comparing with eq. (4.26), leads immediately to identifying d — β/ν with the
similarity dimension,
D = d- β/ν. (4.28)
From table 4.2, we find that D = 91/48 « 1.89 in two dimensions, and D « 2.5 in
three dimensions.
Let us end our discussion of the geometrical properties of the percolation
model with a discussion of lattices of finite length L. What we are aiming at is
known as finite size scaling, which is a powerful way of analyzing data in order to
determine critical exponents. The method is general and by no means restricted to
only geometrical properties. The basic idea is that the system has no other length
For a fractal object D is non-integer.
128 A. Hansen / Disorder

scale than the correlation length £ in addition to the lattice constant and L. Let us
for example analyze the density of the infinite cluster, P ^ , along these lines. We
know the behaviour of this quantity near pc in an infinite lattice,

Poo~(p-Pcf~C0/v- (4-29)

In a finite lattice, as long as the correlation length is shorter than the size of the
lattice, L, eq. (4.29) should be valid: the behaviour of P ^ is determined by what
happens on scales smaller than £, and is therefore insensitive to L. However, if p
is close enough to pc so that ξ is larger than L, P ^ must be a function of L. The
assumption underlying the finite-size scaling method, is that the lattice size L may
be seen as an effective correlation length. This makes sense since the size of the
lattice puts a cut-off on the size of the clusters. With this assumption, we may
write
Pco-ip-Pcf-L-0'", (4.30)
which is the finite size scaling expression for Ρ ^ . To be somewhat more general, we
introduce a function φ(χ), with the asymptotic properties φ(χ) ~ x~®lv for x <C 1
and φ(χ) —> constant as x —► oo. We may then write

Poo(p, L) = (p- pcf<f>{Lli{p)). (4.31)

This expression summarises the contents of eqs. (4.29) and (4.30).


The usefulness of this scaling, is that it provides the most accurate way
to determine the critical exponents. In a numerical simulation one sets p to p c ,
and determines the quantity of interest for several different lattice sizes. The data
obtained are then plotted against L in a log-log plot, and the slope of the straight
line along which the data (hopefully) lie, is the wanted critical exponent multiplied
by —1/u. The difficulty with this approach is that pc must be known to a fairly
high accuracy. If it is not known, one may try to localize pc by trying to produce
as straight a line as possible in the log-log plot. However, this procedure may
be dangerous as a result of the terms marked "less singular terms" in eq. (4.11).
These less singular terms typically also have a power law dependence on p — p c , and
introduce curvature in what ideally should have been a straight line at pc. Thus, a
p that produces an apparent straight line, may be away from the threshold if the
considered system sizes L are not large enough. As a result the critical exponent
one is determining comes out with a - in some cases very - wrong value. As a rule
one should determine critical exponents by more than one method, for example by
using the exponent relations presented above.

4.3 Electrical conductance


We introduced the concept of percolation through the conductance Σ of
a percolating network where the bonds were either conductors or insulators. The
calculation was done with a mean field approximation, which we later argued breaks
3. Electrical conductance 129

down near the percolation threshold. What expression should replace the result of
this calculation, eq. (4.9)? We will answer this question in this section, but we
will also go much further: We will discuss in detail the entire current distribution
near pc. The results we obtain are quite general for scalar transport properties in
disordered materials near a critical point - a situation which will be met in chapter
7 in situations different from percolation.
What replaces the mean field expression for Σ is

Σ = Σ0(ρ-ρ€)\ (4.32)

where t is called the conductance exponent. Pinning down the correct value of the
t exponent has taken several years since it was introduced in 1973^ - after several
large fluctuations, the present value now stands at 1.300 ±0.005 in two dimensions
due to calculations done on a special-purpose computer at Saclay.^ This value,
and the values of t in other dimensions are listed in table 4.1. From eq. (4.9), we
see that the mean field value of this exponent is one, and, even though we have
not shown it, this value is found in all dimensions with this method.
The network we have discussed until now consists of a random mixture
of bonds with zero or unit conductance. What happens if we have a mixture of
bonds with two different values for the conductance? Let us first go to the other
extreme: The conductances are either unit or infinite. That is, we have a conductor-
superconductor mixture.
There is still a percolation threshold in this case. Let p now be the proba-
bility that a given bond is superconducting, and q — 1 — p be the probability that
it has a unit conductance. Then, there is a threshold pc such that for p larger than
Pc, the conductance of the entire network in the thermodynamic limit is infinite.
If p is less than j?c, the conductance of the network is finite, having a power law
dependence on (p — pc) close to pc:

r ib-Pc)-*, if/><Pc; (433)


I oo, if p > pc.
The percolation thresholds for the various lattice types are the same in this case
as in the conductor-insulator network. This can be understood by noting that
to change a given conductor-insulator network into a superconductor-conductor
network we only need to rescale all the conductances: Set the conductances that
are 0 in the first lattice to a very small number e. Then rescale all the conductances
in the network by 1/e, and let e —> 0. The result is that the bonds that before had
a zero conductance now have a unit conductance, and those bonds with a unit
conductance have turned into superconductors. If now the original network was in
a percolating phase, it is now in a superconducting phase, i.e. there is a cluster
of connected superconducting bonds spanning the network from one border to the
opposite one. If we lower p, such that we go below the percolation threshold p c ,
there is no percolating cluster of conducting bonds in the original network. By the
130 A. Hansen / Disorder

rescaling of the conductance argument we have just presented, the new network
cannot have a spanning superconducting cluster. Thus, the percolation thresholds
in the two problems are the same. However, the conductance exponents t and
s are not equal in the two problems (except in two dimensions, as we will later
show). The reason for this is that the Kirchhoff equations that we have to solve in
order to find the currents in the network are not the same in the network before and
after the rescaling of the network. In the insulator-conductor mixture the KirchhofF
equations at a site i with nearest neighbours in is

Σ * Α > ί - Ο = 0, (4.34)
in

where </^n is 0 if the bond between nodes i and in has a zero conductance, and 1 if
it has a unit conductance. After the rescaling of the conductances, 4 the KirchhofF
equations take the form

E ( 1 " Ä . i - ) ( « < - « * » ) = 0. (4.35)


in

with the additional set of equations

9iM ~ vj) = 0, (4.36)

which expresses that the voltage drop across a superconducting bond is zero. It is
the extra set of constraints on the voltages given by eq. (4.36) that results in the
difference between the currents in the statistical properties of the currents in the
original problem and the conductor-superconductor problem.
Let us now study the more general case of having bonds with two types of
conductances, σ\ < σ 2 . ^ The conductance of a network with a fraction p of the
bonds having a conductivity σ2 and a fraction q = 1 — p having a conductance σι,
must obey the following scaling law

Σ(ρ-ρο\σ1,\σ2) = ΧΣ(ρ - ρ€,σ1,σ2). (4.37)

This equation simply expresses that the overall conductance of the network scales
as the conductances of the bonds it consists of. A more subtle scaling law is this
one:
Σ(λ(ρ - P c ) , A 1 + Vi, \ι'*σ2) = ΧΣ(ρ - p c , σ χ , σ 2 ). (4.38)
This equation is constructed to contain both eqs. (4.32) and (4.33). To demonstrate
this, set G\ — 0. Then we have

Σ(ρ - pc, 0, σ 2 ) = λ" 1 Σ(\(ρ - Pe), 0, λ 1 " ' ^ ) . (4.39)


4
This rescaling may be done formally by setting gij = (1 — eij)e + β;^/(1 — e), where e;j is a
random variable taking on the values zero or one. When e — ► 0 we have the insulator-conductor
mixture, and when e - » l w e have the conductor-superconductor mixture.
3. Electrical conductance 131

Now, setting λ = l/(p - p c ), we get

Σ(ρ - ft, 0, σ2) = (p- Pc) 27(1,0, (p - ΡοΥ~1σ2) = 27(1,0,1) σ2 (ρ - pc)*. (4.40)

A completely analogous argument produces the following equation,

Σ{ρ - p c , σχ, oo) = 27(1,1, oo) σλ (ρ - pc)~8. (4.41)

Suppose now that p = pc. Then, we have that

Σ(0,σι,σ2) = (ί/Χ) Σ(0,λ'+1σ1,\1-*σ2)


= \-*σ2Σ{0,\-{'+ί)σ1/σ2,1)
= Σ(0,1,1) σ2 (σι/σ2Υ«>+'Κ (4.42)

This equation shows that for p = pc, the conductivity scales with the ratio between
the conductances of the two components to the power t/(s + t).
The concept of duality transformations has proven to be very powerful,
especially in two-dimensional networks. We now demonstrate how by such a trans-
formation we are able to show that s — t in two dimensions. Our demonstration is
based upon the derivation by StraleyJ10^ This has as a consequence that the scaling
exponent in eq. (4.42) is 1/2.
The Kirchhoff equations for the conductor-insulator network are given in
eq. (4.34). They may be derived through the minimization of a quadratic sum over
voltage differences in the network, (1/2) Σ^ σ^ (vi — Vj)2 where σ^ the conductance
between nodes i and j . The minimum of this sum is the energy dissipation in the
network, W,
W = mm\Yjaij{vi-vj)\ (4.43)

and determines the voltages that solve the Kirchhoff equations, Σ · σ^ (vi — Vj) — 0.

Let us now work with the currents in the bonds of the network rather than
the voltages at the nodes. We call i^ the current flowing through bond ij. This
current may be decomposed into a sum over loop currents. A loop is the closed loop
of bonds that form the borders of a unit cell in the network. An example is shown
infig.4.6. The loop current is the current flowing in the loop forming the loop.
Infig.4.7 we show a bond ij in a square lattice. This bond forms one side of two
loops m and n. The current flowing in this bond, i^ may be decomposed into the
two loop currents zm and in: ijj = im — in. If we call the resistance of the bond i j ,
rmn = 1/σο? w e m a Y write the dissipation in this bond as (l/2)r m n (z m — in)2. We
may now define a quadratic form by summing this expression over all loops in the
network: (1/2) Y^mn rmn(im — in)2. The minimum of this expression with respect to
the loop currents is again equal to the dissipation in the network

W = min ! Σ rmn{im-inf. (4.44)


K
mn
132 A. Hansen / Disorder

Figure 4.6 A loop k in a square lattice. The loop current ύ is the current flowing in it.

m n

Figure 4.7 The current i^ flowing through bond ij decomposed into the two loop currents in
and in; iij = im — in. We have also indicated the dual lattice by broken lines.

Thus, the set of loop currents zm that minimizes the right hand side of this equation,
solves the current distribution in the network, and forms an alternative formulation
of the Kirchhoff equations.
It should be noted here that the dissipation W is equal to the resistance R
of the network if the current flowing through the network is unity.
3. Electrical conductance 133

We now introduce the dual lattice. We construct this by replacing each loop
in the original lattice by a node, and joining these nodes by bonds that cross the
bonds in the original lattice. On a two-dimensional triangular lattice the dual is a
honeycomb lattice, on a three-dimensional cubic lattice the dual is a face-centered
cubic lattice. The original and the dual lattice are in general different - except for
the two-dimensional square lattice. This lattice is self-dual.
One consequence of this self-duality in the square lattice, is that eqs. (4.43)
and (4.44) have exactly the same form. If we now interprete the loop current im in
the original lattice as a node voltage vm in the dual lattice, and set the resistance
of bond rmn = Ι/σ^ equal to the conductance, σ| ηη , of the bond ran in the dual
lattice eq. (4.43) takes the form

WD = nnn±j:a'mn(vm-vn)2,
v
(4.45)
k ^ n
ran
where WO is the dissipation in the dual network. As above, we notice that the
dissipation in the dual network, WD, is equal to the conductance ΣΌ of this network
if the voltage drop across the network is unity. We also notice that the total current
flowing through the original network transforms into the voltage drop across the
dual network.
We now return to the original problem of having a binary mixture of con-
ductances being either unity (with probability p) or zero. On the dual lattice this
has been transformed into a binary mixture of conductors (now with probability
p) and superconductors (with probability q = 1 — p).
In the original network, the conductance goes to zero as Σ ~ (p — pc)1.
Thus, the resistance of this network diverges as (p — pc)~*· Through the duality
transformation, this resistance R is equal to the conductance, Z D , of the dual net-
work which is a conductor-superconductor mixture. This follows from eqs. (4.43)
and (4.44) being equal. Near qc (since p —> q and q —> p in the duality transforma-
tion) the conductance of the dual lattice diverges with an exponent s. This leads
to the conclusion that s = t on the two-dimensional square lattice, and since these
exponents are only dependent on the dimensionality of the networks, we have that
s = t in two dimensions.
It is now trivial to show that the percolation threshold in the two-dimen-
sional square lattice is 1/2. The probability to have a conducting bond in the
original lattice is p, while the probability to have a superconducting bond in the
dual lattice is q = 1 - p. From the previous discussion of the relation between the
conductor-insulator problem and the conductor-superconductor problem, we know
that qc = 1 — pc = pc since the network is self dual. This leads to the anticipated
result.
Until now we have only discussed one aspect of the transport properties
near the percolation threshold, namely the behaviour of the conductance. We now
enlarge this by studying the entire current distribution in this region. This will lead
into the subject of multifractality, which we introduce and discuss in rather general
134 A. Hansen / Disorder

Figure 4.8 The current-carrying backbone of a conductor-insulator network at the percolation


threshold.

terms. The reader should keep in mind that these methods and ideas are far more
general than the percolation problem itself. In the following we will specifically
have in mind the conductor-insulator problem.
The central quantity in our subsequent discussion, is the current distribu-
tion - or if one prefers, current histogram - n(i,p). Let us now set p = pc, and work
with finite-sized networks of finite size. By using the finite-size scaling arguments,
we will generalize our results to the case of p ^ pc. Thus, we work with n(z, L).
Our first question is where does the current go within the network at pc.
In fig. 4.8 we show an example of the current path in such a network. This path is
usually referred to as the backbone. In this case, a network of size 256 x 256 was
generated, then a current was injected at the lower left side of the network, and
extracted at the upper right corner. Let us call these injection points - and the
bus bars used earlier in this chapter - "electrodes". Readers with a magnifying
glass and a skeptical mind will protest that the backbone shown in fig. 4.8 contains
bonds that carry no current. These bonds are the perfectly balanced ones, as the one
shown in fig. 4.9. However, if there are some variations, no matter how weak, in the
conductances of the present bonds, the perfectly balanced bonds will start carrying
3. Electrical conductance 135

—►
m
1—► ' **—'—^—

Figure 4.9 The bond marked A is perfectly balanced, and does not carry any current. It is
therefore not part of the current-carrying backbone, but part of the geometrical backbone.

a current. In reality, the backbone shown infig.4.8 is a geometrical backbone, while


the backbone with the perfectly balanced bonds removed is called the current-
carrying backbone. The geometrical backbone has its name since it is determined
by purely geometrical means: Any bond belonging to it has two non-intersecting
paths which connect one end of the bond to one electrode, and the other end
to the other electrode. It turns out that the density of perfectly balanced bonds
in the geometrical backbone is a constant: For the square lattice with point-like
electrodes, as shown infig.4.8, about 0.5% of the bonds are perfectly balanced,
and in the bus-bar geometry, about 0.15% are perfectly balanced. These numbers
are independent of the size of the networkJ11»12!
The backbone is a subset of the infinite cluster. Those bonds in the infinite
cluster not belonging to the backbone constitute the dangling ends for obvious
reasons. Measuring the fractal dimension of the backbone leads to a surprise: It is
quite different from the fractal dimension of the infinite cluster itself. The fractal
dimension of the backbone, DB, is in two dimensions 1.62 ± 0.02j13^ which should
be compared to the fractal dimension of the infinite cluster being D = 1.89. No
relation between these two numbers is known.
The mass of the current-carrying backbone is the zeroth moment of the
current distribution n(i,L),

Σ i°= £ n(i,L)~LD*. (4.46)


bonds ij currents i

Returning tofig.4.8, we see that a fair number of bonds carries all the
current flowing through the network. These bonds are known as links or cutting
bond, since by cutting one of these, the network stops conducting. It has been
shown rigororously by Conigliof14^ that the number of these bonds scales with the
network size as \jv where v is the correlation length exponent defined in eq. (4.15).
Thus, the density of these bonds on the backbone decreases with increasing lattice
size, in two dimensions \jv — 3/4.
Those bonds belonging to the backbone which are not cutting bonds consti-
tute the blobs. Skal and ShklovskiiJ15! de GennesJ16^ and S t a n l e y ^ have suggested
a very useful picture, known as the nodes-links-blobs picture, for the backbone in
a percolating network with a correlation length ξ. At length scales larger than ξ
the backbone is homogeneous. At length scales less than ξ it is fractal. Let us now
136 A. Hansen / Disorder

Figure 4.10 An illustration of the nodes-links-blob picture of the backbone.

imagine a "super network" with lattice constant £. The "bonds" in this "super
network" are structures as the one shown in fig. 4.8, consisting of links and blobs.
The resulting network is shown in fig. 4.10.
We will not pursue the discussion of the geometrical properties of the back-
bone further, even though there are a significant number of other properties known
and surprising results. A suitable place to dig further into this subject is Stanley's
lectures from the 1986 Cargese Summer SchoolJ18^
The number of bonds belonging to the backbone is related to the zeroth
moment of the current distribution n(z,L). The conductivity is related to the sec-
ond moment. To see this, we refer back to eq. (4.43). If the set of currents iij
solves the Kirchhoff laws, the dissipation in the network with conductance σ^ of
the bonds equal to unity, is given by

W = E(AVf = ! £ $ · (4.47)
u
If we set the voltage difference across the network to unity, we see that the con-
ductance is directly the second moment of the current distribution,

27 = £ t 2 n v ( i , L ) . (4.48)
i

The subscript V on the distribution function nv(i,L) indicates that the voltage
difference across the network has been held fixed (and equal to unity). This is
an important piece of information when averaging over an ensemble of networks
- a fact we will discuss in a moment. We call this ensemble the constant-volt age
ensemble. By appealing to eq. (4.32), and the finite-size scaling argument, we find
3. Electrical conductance 137

that the second moment of the current distribution scales as

(i*)v = £ i2nv(h L) ~ L~t,v. (4.49)


i

We also know another moment from the preceding discussion. We stated


above that the number of cutting bonds scales as L1^. If the current flowing
through the network is unity, we know that the cutting bonds must also carry a
unit current, since by definition they carry all the current that flows through the
network.
If we now calculate the fcth moment of the current distribution with a unit
current flowing into the network, and let fc go to infinity, we find5

lim (i*) 7 = lim Y t*n 7 (i, L) + n 7 (l, L) ~ L1/u, (4.50)


k—>oo k—KX) 7~~i

since the first term in the middle expression above goes to zero as fc —> oo. The
subscript I in this equation indicates that we are in a constant-current ensemble.
We have now introduced the constant-voltage and the constant current
ensemble. The concept of different ensembles is important. The reason for this is
that we need to keep some quantity fixed in order to compare the different samples
(i.e. realizations or networks) that we average over. The quantity we choose to
keep fixed defines the ensemble. The scaling exponents we find are different in the
different ensembles, but they are related to each other.
To formalize this discussion more, we call R a given realization out of a
total number of NR. The fcth moment of the current distribution is then

iV
* R JR

where ijR is the current through bond JR in realization R. Let us choose the currents
ijR to be those of the const ant-volt age ensemble, where the voltage difference AV
across each sample R is equal to unity. Let us now define a scale factor SR that we
use to rescale the const ant-volt age currents ijR: ijR —> SRIJR. This defines a new
ensemble, s, and the corresponding fcth moment is

N
* R 3R

In particular, the constant-current ensemble, where a unit current flows through


each realization, is found by setting sR = l / Σ β - the inverse of the conductance
5
This limit is subtle. As we do it here, letting k go to infinity, and then deducing the scaling
law for large L, may not give the asymptotic behaviour of the large moments of the current
distribution. However, going to the thermodynamic limit, L —> oo, first, and then let k —> oo, will
correctly reproduce this behaviour. That switching these two limits may lead to different results,
is caused by the cutting bonds that govern the limiting behaviour of eq. (4.50) form a vanishing
subset of the backbone.^ 11 !
138 A. Hansen / Disorder

of realization R. The reason for this is simply the relation I = ZRAV. Setting I
- the current flowing into the network - equal to one demands that the voltage
difference across the network is reduced from unity to Ι/Σ^ which in turn reduces
all currents within the network by the same factor. This leads to the relation

(i*>, = (Σ-" ik)v (4.53)

between these two ensembles. If the fcth moment is self-averaging, meaning that
the relative fluctuations vanish as the lattice size increases, we may decompose the
right-hand side of eqs. (4.51) and (4.52) into

<**>. = (sk ik)v = <«*M»*>v, (4-54)


and, specifically for the constant-current ensemble,

{t*)/ = 27-*<**>v ~ Lki (ik)v. (4.55)

From eqs. (4.46) and (4.55), we see that the fractal dimension of the backbone is
the same in all ensembles. Using eqs. (4.49) and (4.55), we find that the second
moment of the current distribution in the constant-current ensemble behave as

(*2>/ = Σ < 2 η / ( ί ^ ) ~ ^ / " · (4-56)


i

We calculated the infinite moment, eq. (4.50), in the constant-current ensemble. By


eq. (4.55), we see that the corresponding exponent in the constant-voltage ensemble
does not exist.
How do the other moments behave? Rammal et al^ and de Arcangelis
et aÜ2°} essentially simultaneously demonstrated that the other positive moments
scale as power laws in the lattice size,

(i*)j~Lw(fc), (4.57)

and that these exponents fall on a curve which is anything but a straight line when
plotted against k. This is shown in fig. 4.11. The fact that these exponents do not
have the form ak + 6, where a and b are some constants, is significant. It signals
that we are dealing with a multifractal phenomenon. The opposite situation, is
called constant-gap scaling. We will now analyze the significance of multifractality.

The starting point of our analysis - based on Halsey et al^ and which is
known as / — a formalism - is 6
1
(ik)j = Σ ikni{i, L) « Jdi ikm(i, L) ~ Lyi{h). (4.58)
« o
6
The reader aquainted with statistical mechanics will discover that this is identical to the
derivation of thermodynamics from the partition function.
3. Electrical conductance 139

y(k)

Figure 4.11 The scaling exponents of the fcth current moment in the constant-current ensemble
plotted against k.

If this moment is self-averaging, the integrand iknj{i, L) has a peak somewhere


which becomes sharper with increasing L. With this assumption (which has been
verified numerically) we may estimate the integral in eq. (4.58) by using the saddle
point method.
The idea of this method is to rewrite the integral as
1 1

j di i*n/(i,L) = l di L·k log(i)/log(L)+/(t,L) (4.59)


o o
where we have defined the function
log(nj(i,L))
f(hL) = (4.60)
log(L)
We also introduce a different variable for the current,

log(i)
<*i(i,L) = (4.61)
log(L)'

where the subscript I indicates the ensemble.7 When the integrand has a very
pronounced peak, the integral gets its main contribution from that point. The
7
We do not use a subscript on the function / to indicate the ensemble, because it is invariant,
as we will demonstrate later.
140 A. Hansen / Disorder

position of this peak is found by determining where the first derivative of the
exponent in the integrand, is zero. This is equivalent to

!ψ&-*. («2)
The value of a that solves this equation is called a(fc,L). Thus, the integral of
eqs. (4.58) and (4.59) becomes
1
[ di £,-fca/+/(a/»L) — lj-kaI{k,L)+f{otI{k,L),L) ^ jj)i{k) (4.63)
0

We identify
yi(k) = -fca,(fc, L) + /(a 7 (fc), L). (4.64)
Taking the derivative of this equation with respect to k gives
άΜ
dfc
= -al{k,L)
'
- \fd(Wft,L)-/(aKfc^)))\
da/(fc,L)
ί^φΒ]
) \ Ak = _0/(ft>L).
(4.65)
The second equality comes from using eq. (4.62). Since the left-hand side of this
equation is independent of /, so must the right-hand side,

a/(fc,L) = a/(fc). (4.66)

Returning to eq. (4.64), moving the first term on the right-hand side over to the
other side of the equality, we conclude that also the function / must be independent
of L:
/(a/,L) = /(aJ). (4.67)
The reader should notice carefully the structure of this web of mathematical
manipulations. We have two functions f(oti) and yi(k) of the two variables a/ - a
rescaled current - and k. Through eq. (4.62), we see that the variable k is conjugate
to a/, and the relation eq. (4.64) is nothing but a Legendre transformation from
the variable aj to the variable k. Thus, the two functions /((*/) and yi(k) are
Legendre transforms of each other.
One important lesson from these manipulations is to identify what the
thermodynamically intensive variables in this problem are, namely the rescaled
current distribution, eq. (4.60), the rescaled current, eq. (4.61), the rescaled fcth
moment, yi(k) = log((z*)/)/log(L), and of course fc.
But, where has the multifractality gone among all these mathematical ma-
nipulations? It is hidden in the shape of the rescaled current distribution function
/ . To see this, assume that we are dealing with a phenomenon showing constant-
gap scaling: y(k) = ak + b. By using eq. (4.64), we identify a = a and f = b,
thus both being constants. Thus, a plot of / versus a comes out to be but a single
3. Electrical conductance 141

1 1 1 1 1 1 Ί 1 1 1
+
+ -
±, H iJ if rh rfi. i

^ * * * % ;
Γ jutW&l·
^ ^

f- xz
■*

Γ ^*%{r ♦
1 \— t _
o
+
-Φ-
1 + J ^
0 L— V L=64
on
r^t
F#
+
p- _
+


1
1
1

,_, , u
1

-2
1

-2 -1 0
-a=log(i)/log(L)

Figure 4.12 The multiiractal spectrum of the insulator-conductor mixture at the percolation
threshold in the constant-current ensemble.

point. Multifractality is then the situation when the function / ( a ) is more than a
point. In fig. 4.12 we show the shape of this function - also known as a multifractal
spectrum - for the current distribution in the constant-current ensemble.
The word "multifractality" implies that this phenomenon is an extension
of the notion of a fractal. However, up to this point, all it has signified is that
the various moments of the current distribution scale with different non-trivial
exponents, or equivalently, that the distribution is "broad". The basis for our anal-
ysis was that each moment of the current distribution gets its main contribution
from a precise part of the distribution function. Suppose we pick the fcth moment.
142 A. Hansen / Disorder

Inverting eq. (4.61), with a/ equal to a/(fc) gives

i(k) ~ L~ai{k). (4.68)

The interpretation of this equation is simple. It states that the current i(k) that
dominates the fcth moment, scales with an exponent — aj(k). Then, using eq. (4.63),
we may write
(ik)i~i(k)k Lf{ai{k)). (4.69)
We are now able to recognize /(a/(fc)) as the fractal dimension of the subset of cur-
rents i(k). In particular, going back to eq. (4.46), we see that /(a 7 (fc = 0)) = D^.
This justifies the choice of the name "multifractal." The multifractal may be seen
as an interwoven set of fractals, with different fractal dimensions. It is important
to notice, however, that there must be a way of distinguishing these various fractal
subsets - they must be tagged in some way. In the case of the current distribution
this tagging was possible through the above grouping of currents. But, for example
in the case of the geometrical properties of the percolation clusters that we dis-
cussed in section 4.2, no such tagging is possible, since all bonds have but one type
of property; present or absent.
How does a change in ensemble affect the multifractal spectrum / versus
a/? We return to eq. (4.54). Furthermore, assume that {s)j ~ Lx. If then (ik)8 ~
Lya{k\ we have that
y,(k) = yi{k) + kx, (4.70)
which, by using eq. (4.64), leads to

kas(k) - f(a.(k)) = k(aj(k) + x) - f(a.(k)). (4.71)

If we now set
as = a! + x, (4.72)
and note that the change from one ensemble to another only affects the overall
scale of the currents in each sample, and not the relative distribution of currents,
we conclude that the fractal dimension of the various sets of currents cannot change.
Thus, a change in ensemble is nothing but a translation of / along the a-axis.
The reason for putting so much emphasis on the question of ensembles is
that even though in the thermodynamic limit, L —> oo, they are all equivalent,
for finite sizes, the fluctuations of the quantities one is measuring may be much
larger in one ensemble than in another. An example is the fourth moment of the
current distribution in the constant-voltage and constant-current ensemble. From
eq. (4.53) we see that the fluctuations will be much smaller in the first of these
two ensembles, since the fluctuations in the currents will be compensated by the
fluctuations in the conductance: a large fourth moment for a given realization also
results in a large conductivity.
This effect becomes dramatic when we discuss the negative moments of the
current distribution. Neither in the constant-current nor in the constant-voltage
3. Electrical conductance 143

ensembles, the negative moments are self-averaging. Ensembles may, however, be


defined where they are self-averagingJ12^ Let us for sR - defined in eq. (4.52) -
choose

**=(ΣΓ2) · (4·73)
JR

The point of the above choice is that this quantity fluctuates as badly as the small
currents that govern the negative moments, and by forming an ensemble with this
normalization, these fluctuations are compensated for. In this ensemble - called the
a
—2"-ensemble - the minus second moment is constant and equal to one. When
numerical simulations were done in this ensemble, it was found that a_ 2 ( _ &) ~ 0,
and / ( α _ 2 ) ~ 0, which signals constant-gap scaling: All the negative moments are
governed by the same set of currents. The number of bonds in this set grows slower
than a power law in L since the fractal dimension is seemingly zero.
Is there any way of relating the exponents £/-2(&) in this ensemble and
those of the constant-current or constant-voltage ensembles, through relations sim-
ilar to eqs. (4.70) and (4.72)? It does not seem to be possible, since the positive
moments are not self-averaging in the "—2"-ensemble. The situation is then as
follows: Through the const ant-volt age or constant-current ensembles we have been
able to study the large-current side of the multifractal spectrum. However, as can
be seen in fig. 4.12, the part of the spectrum to the left of the maximum, which
is governed by the negative moments, there are large finite-size effects. This is a
reflection of the lack of self-averaging of the negative moments in these ensembles.
On the other hand, by using the "—2"-ensemble, we have been able to focus on the
side of the spectrum that we could not study through the previous ensembles. It
is, however, not possible to study the entire spectrum at the same time. How far
apart is the right-hand endpoint of the curve from the left-hand endpoint? We can
only guess, but it is now generally agreed that the two endpoints are infinitely far
apart. This has as a consequence that the left-hand side of the spectrum becomes
parallel with the α-axis. The argument behind this assumption is that it is possible
to find configurations containing currents that falls off in size exponentially as L
increases J22^
It is amusing to note that the bonds carrying these smallest non-zero cur-
rents are very different from the perfectly balanced bonds that we discussed earlier.
This may be seen through noting that the fractal dimension of the the perfectly
balanced bonds is equal to the fractal dimension of the backbone itceself, DB, but
the fractal dimension of the smallest non-zero currents is zero.
The situation in the random conductor-superconductor network is com-
pletely analogous to the one we have just described: There is multifractality in
the current distribution. In two-dimensional networks, the multifractal spectrum
is identical to the one found in the random conductor-insulator network. This is a
consequence of the self-duality of the square lattice discussed earlier in this chapter.
However, the scaling exponents of the current moments are related in such a way
144 A. Hansen / Disorder

that the const ant-volt age ensemble moments scale in one system as the constant-
current ensemble moments scale in the other, and vice versa, de Arcangelis et a/'23'
have discussed the multifractal spectrum of a mixture of bonds with small (σ χ )
and large conductances (σ2) at the percolation threshold for the bonds with small
conductances - thus a generalization of the result of Straley^1^ that the scaling
exponent in eq. (4.42) is 1/2. They found that the kth. moment scales as

<«*/»>, ~ ( g ) , (4.74)
where e is the power dissipated in the bonds - this being the natural generalization
of the current moments for these mixtures.
So far our discussion of the multifractal properties has been done at the
critical point pc. A natural question is to ask what effect does multifractality have
away from the critical point pc. This question has been discussed by Roux et al
t24^ by using finite-size scaling arguments along the lines described at the end of
section 4.2. We summarise some of their discussion as an example of finite-scaling
analysis.
Instead of having a lattice size L being the only length in the problem
(besides the lattice constant), we now also have the correlation length £ ~ | p —
Pc \~v\ eq. (4.15). Suppose that we are at p = 1. Then the network is homogeneous,
and all currents in it are equal and scale as Ll~d in the constant-current ensemble.
The number of bonds carrying this current, scales as the volume of the network,
n(z,L,p = 1) ~ Ld. Thus, we have f° = d and a°T = d — 1. The superscript 0
means that these are values for the homogeneous network. Knowing this, and the
current distribution at the percolation threshold, p = pc, we may reconstruct the
distribution for £ < L,

ni(i,L,p) = {LKfnI{i(Ltt)a'i,t,pe). (4.75)

This equation is easy to understand in terms of the nodes-links-blobs picture of


refs. [15,16,17]. The distribution n/(z(Z//£) a ', £,p c ) contains the current distribution
flowing into one "superbond" of length £, and the prefactor (L/ξΥ counts the
number of these superbonds in the entire network. The current distribution in one
superbond is by definition of the constant-current ensemble given for a unit current
flowing through it. Thus, the currents flowing through each bond within a given
superbond i must be scaled to correspond to a unit current flowing through the
superbond. Since the network is homogeneous at the level of the superbonds, a
current (£/L)~ai flows through each superbond. Thus, we scale each current inside
a given superbond down by a factor (£/L)~ai.
Now, introducing the rescaled quantities of eqs. (4.60) and (4.61) for the
current distribution in the entire network, and for the current distribution in one
superlink, we get

log(ni(i,L,p))/log(L) = f + (log(0/log(L))(/(a,) - / ° ) , (4.76)


4. Elasticity with angular stiffness 145

and
- log(i)/ log(L) = a? + (log(i)/ log(L))(a, - a?), (4.77)
where
a^-log^/L^/log«) (4.78)
is the thermodynamically intensive variable for the currents within one superbond,
and where f(oti) is the corresponding thermodynamically intensive variable for the
current distribution in this superbond. Equations (4.76) and (4.77) express how the
current spectrum looks away from the percolation threshold. Closer examination
of them reveals that they are equations for a dilation of center (α°, / ° ) , and ratio
log(£)/log(L) of the function / ( a 7 ) .
We have now finished our survey of the electrical - or in general "scalar" -
transport properties in a percolation context. However, there are other "transport"
problems that occur in a percolation-type connection, which are also technologically
very important - besides being of genuine scientific interest. We therefore continue
with reviewing what is known as "elastic percolation".

4.4 Elasticity with angular stiffness


Suppose now that our network of bonds, present with probability p, or
absent with probability q = 1 — p, is an elastic network. What happens in this case
when we get close to the percolation threshold pcl
It turns out that there are two classes of elastic network models that may
be defined, depending on what kind of elastic properties the individual bonds
have. The first class consists of bonds that transmit all the elastic information.
The second class transmits only partial elastic informationJ25^ We will discuss the
implications of this in the next section.
In this section we will concentrate on the first of these two classes of elastic
models. The first percolation model of this type to be introduced was the "bond-
bending" model of Kantor et alS26^ If we let Uj be the displacement of node z, y^
be 1 if bond ij is present and zero otherwise, e^ be a unit vector in the direction
of bond ij starting at node z, θ^ be the angle between bonds ij and zfc, and β
and 7 be two elastic constants, we may write a function in analogy with eq. (4.43)
whose minimum gives the elastic energy of the network,8

H = \βΣff«((«i - *i) ■ e o ) 2 + h Σ 9ij9ik^ik- (4-79)


ij jik
8
A simplified expression, known as the Born model}27} has the following form

ij ij

It is simpler than the bond-bending model to deal with, both analytically and numerically. How-
ever, it suflfers from the serious drawback that it is not a representation of an elastic problem, as
the above expression is not rotationally invariant.
146 A. Hansen / Disorder

This function is not a quadratic form, giving rise to linear elastic equations. The
reason for this, is the unit vector e^ that appears in the first factor. This vector
points along the bond, and as the bond is displaced, so is e^. However, the equations
can be linearised by taking this unit vector to be the one of the undistorted lattice.
The minimum of eq. (4.79) gives the elastic energy of a network of Hookean springs
with bond-bending terms.
A different model, falling into the same class as the bond-bending model, is
the beam lattice of Roux et alS28^ This model was described in detail in chapter 3.
The equivalents of the conductance, 27, in electrical transport, are the
macroscopic elastic moduli of the network, E (Young modulus) and G (shear mod-
ulus). It is clear that if the probability of having bonds present drops below the per-
colation threshold, the elastic moduli will be zero just as the conductance because
the network will not be connected. Above the percolation threshold the network
is rigid, ensured by the existence of at least one connected path throughout the
network. This is caused by the angular forces. The percolation threshold is thus
also a threshold for rigidity in these elastic models, and we have

E~G~{°' .Γ * * < * ' (4.80)


\\P-Pc) , ifp ^Pe-
in section 4.3, we also introduced a conductor-superconductor percolation problem
in addition to the conductor-insulator one. We can do the same here. Until now,
we have defined the rigid-non-rigid percolation problem. The equivalent of the
conductor-superconductor problem, is the rigid-super-rigid percolation problem.
Here we have a mixture of a fraction q — 1 — p of bonds with unit elastic moduli,
and a fraction p of bonds with infinite elastic moduli - thus being infinitely rigid.
The equivalent of eq. (4.80) in this case is
5 if
E~G~ f(p-P^ ' P<i>c; (4.81)
\oo, Hp>Pc-

Through the self-duality of the two-dimensional square lattice we were able


to show that s = t in the electrical conduction case - s and t are defined in
eqs. (4.32) and (4.33). A similar duality argument does not work here, and in fact
S φ Τ in two dimensions. The duality transformation is, however, capable of giving
valuable insight into a related elastic problem; namely that of out-of-plane flexion.
We find, as we will demonstrate in the next section, that the in-plane elastic moduli
transform into flexural compliances, if, and vice versa. This result combined with
the singular behaviour of the elastic moduli near the percolation threshold, leads
tot29!

*~ί?' ^ ΐΡίΑ5 (4·82)


{(P-Pe) , UP>Pc,
in the case of the rigid-super-rigid mixture, and in the rigid-non-rigid mixture, the
4. Elasticity with angular stiffness 147

Figure 4.13 We apply a force F to the backbone at the percolation threshold. The torque felt
at node i is Fr t , and the force is F. We may also apply a torque M — F£ to the backbone at the
endpoints.

flexural compliance behaves as

K~{(p-p^s> tfjxpc; (483)


[ oo, if p > pc.
The next point that we will discuss, is the relation between the exponents
T and t. The inequality™
T < t + 2v (4.84)
has been proved. However, numerical evidence strongly suggests that this inequality
in reality is an equality: Zabolitzky et al^ have determined T to be 3.96 ± 0.04
in two dimensions. With the result of ref. [9], t = 1.300 ± 0.005, and remembering
that v = 4/3, we find t + 2u = 3.967 ± 0.005.
The argument leading to eq. (4.84) consists of two main simplifications: (1)
The "transport" of torque dominates compared to the "transport" of forces in the
elastic networks near the percolation threshold, and (2) in loopless structures each
component of the torque behaves as a scalar quantity.
Let us for simplicity assume a two-dimensional problem. Our plan is to
demonstrate (1) by finding a lower and an upper bound on the compressional com-
pliance of a percolation backbone, and show that both bounds are both dominated
by the internal torques created by applying an external force. We base our discus-
sion on the model of Kantor et al^ eq. (4.79). The bounds are found by appealing
to the picture of the backbone at the percolation threshold consisting of links and
blobs (remember fig. 4.8). We get a lower bound on the compliance by assuming
that the blobs are infinitely rigid. Thus, let us only count the cutting bonds, whose
number, iV), scale as ξ1/"}1^ Thus, ignoring the blobs, the backbone has a one-
dimensional topology. Let us now compress this one-dimensional structure with a
force F, as shown in fig. 4.13. We concentrate on the bond between nodes i and
j . The compression of this bond is given by U{ — Uj = (l//3)Fcos0^, where φ^ is
the angle the bond forms with the axis between the two endpoints of the chain,
and F =| F |. The rotation δφ^ it has undergone is due to the torque created by
F, δφ^ — (l//y)Fri where Τ{ is the distance between node i and the axis between
148 A. Hansen / Disorder

the two endpoints of the chain. Thus, the lower bound on the total elastic energy,
E^ of the backbone is

(4.85)

where S,2 = (1/JV,) Scutting bonds i · A lower bound on the compliance of the back-
r
bone is to only take into account the bonds belonging to one of the shortest paths
between the two endpoints. We find exactly the same expression as in eq. (4.85),
but with Ni substituted by Nsp, the number of bonds belonging to the shortest
path, 9 and 5 s p substituting for Si. K a n t o r ^ has shown numerically that both S/
and 5 s p scale as £. Thus, as ξ diverges near the percolation threshold, the second
term in the elastic energy in eq. (4.85) will dominate due to the internal torques
in the backbone, both for the upper and lower bound.
The next point, (2), we want to demonstrate is that the torques behave as
electrical current along a loopless chain. If we set β = 0 in eq. (4.79), we see that
the expression we end up with is identical to that of a chain of conductors if we
interpret the angle θ^ as a voltage difference between nodes ji and ik. However,
if the network contains loops, such an identification is no longer possible. This
is due to forces appearing in the loops. These loops lower the actual "amount of
torque" that flows in the network compared to that one would get from solving
the Kirchhoff laws. As shown in fig. 4.13, if we apply a torque M = F£ to the
backbone, this elastic energy becomes

(4.86)

where Σ is the electrical conductivity of the network. This leads to

(4.87)

which leads directly to eq. (4.84).


Let us now go on to describe the second class of percolation behaviour of
elastic networks.

4.5 Elasticity of central-force systems


The percolation properties of the randomly diluted central-force network
was first studied by Feng et alS3^ The model consists of a network of central-
force Hookean springs, and may be obtained from eq. (4.79) by setting 7 = 0.
The behaviour of a diluted network of such springs is very different from that of
the bond-bending percolation networks. This can for example be seen in that this
network is not rigid for p = 1 - all bonds present - if for example the network is
This number scales as ξ 1 · 1 2 9 in two dimensions and ξ 1 · 3 3 8 in threeJ 32 ^
5. Elasticity of central-force systems 149

Figure 4.14 A rigid backbone at pT in a triangular lattice of size 80 x 80. (Prom ref. [36].)

a square, cubic or hypercubic lattice. However, in other lattice types, such as the
triangular lattice in two dimensions, there is a non-trivial rigidity threshold.
The reason for this difference is that a bond may only transmit the com-
ponent of a force parallel to itself. That means, in order to determine whether a
given node is rigidly connected to the rest of the network, it is not enough to check
only its elastic connection through one single bond. That will only tell us whether
it is rigidly connected in the direction parallel to this bond. Thus, we need also to
check the rigidity along some other bonds in directions different from the first one.
This boils down to the rigidity concept in this network being a non-local o n e : ^
in order for two bonds in a central-force elastic network to be rigidly connected
to each other, it is a necessary, but not sufficient condition that there exists more
than one independent connected path between them. To find a sufficient condition
is a very difficult problem that has recently been solved by Crapo, see for example
Guyon et alS35^ There is, however, still no simple way of identifying the bonds that
ensure a rigid connection between two nodes.
This non-locality has for example the effect that there is no equivalent
of the cutting bonds in the rigid backbone - in fig. 4.14 we show an example
of such a backbone in a triangular lattice. It also has the consequence that the
rigidity threshold, pT is larger in this problem than in the connectivity percolation
problem. The rigidity threshold pT the threshold below which there will be no rigid
connection between nodes infinitely far away, while above there will be a finite
probability to find rigidly connected nodes infinitely far apart. This definition is
parallel to that of the percolation threshold, with the concept of "connectivity"
replaced by "rigidity."
150 A. Hansen / Disorder

As mentioned above, there exists a simple geometrical method to extract


the rigid backbone for this problem. There is also no analytic result that leads
to an exact pinning down of pT10 This has led to large difficulties in studying the
critical properties near p r , and at this moment, the quantitative results on the
triangular lattice that exist are controversial. This author, however, will present
his belief of what the Truth is, but the reader should keep in mind that there are
other competing Truths around. [25 ' 34 ' 39-411
There is no controversy in the existence of a critical rigidity threshold
different from p c , the connectivity percolation threshold. As in the angular elasticity
case, the elastic moduli are non-analytic near the rigidity threshold,

s~c~{?' v ^PiP'; (4·88)


We may also define a corresponding rigid-super-rigid network, as was done for the
bond-bending problem. In this case, we have that

25~σ~/(Ρ-Λ)"'· Η*5*; (4.89)


loo, ιιρ>Ρν
In the previous section we claimed that a duality transformation relates the
in-plane elastic behaviour to an out-of-plane flexion problem. We demonstrate now
that this is so for the central-force problem,^ leading to the flexural compliances
behaving as
'oo, ifp<pr;
{
{P-Pr)f, ifp >Pr,
in the case of the rigid-super-rigid mixture of loops, and in the rigid-non-rigid
mixture
F^fiP-Pr)-9, ifP<Pr-, (491)
[
10, ifp>pr, '
in analogy to eqs. (4.82) and (4.83) for the bond-bending case.
This derivation follows closely the one of StraleyJ10] The equivalent of
the loop, and the loop current defined in the scalar case, see fig. 4.6, is now a
"cartwheel" as shown in fig. 4.15. In this construction the force φ is balanced at
each node, and therefore plays the same role as the loop current im of eq. (4.44).
We may now decompose the force fa in each bond ij, of the network into a sum
of "cartwheel forces" φιη. The elastic energy of the network is then given by

Ee =
"£ί η ί Σ ^ ί Ν ύ ' ) + Ψη(χί) ~ ΨΗ(Χ3) ~ Vl{ij)Yi (4·92)
ij

where m(ij)... l(ij) are defined in fig. 4.16. ßij is the elastic modulus of bond ij.
l
0 With one exception, however. In references [37,38] a rigid-super-rigid triangular network was
studied, where all bonds in one direction were super-rigid, while in the other two directions there
was a mixture of rigid and super-rigid bonds. In this case pT = 1/2.
5. Elasticity of central-force systems 151

Figure 4.15 The "cartwheel" which is the equivalent of the loop of fig. 4.6. The force φ is
balanced at each central node in this construction, m is the coordinate of this "cartwheel".

Figure 4.16 An illustration of how the force fa through bond ij may be decomposed into four
"cartwheel forces", m(ij), n(ij), k(ij), and l{ij).
152 A. Hansen / Disorder

Δ7
Figure 4.17 A diamond-shaped elastic tile with aflexuralcompliance /3tJ.

We have now formulated the central-force problem in such a way that we


can define the duality transformation: Let us map the initial network onto a plane
covered with diamond-shaped tiles, as shown in fig. 4.17. For each bond ij with
an elastic modulus ßij in the original network we place a tile ij with a flexural
compliance β^. The flexural elastic energy of this tile network is

EE = min \ Σ&Λζ™(ν) + zn(ij) ~ zK%j) ~ zi(ij))2, (4-93)


ij

where zm^j) is the perpendicular displacement of the node m(ij). Comparing this
expression with eq. (4.92), we see that they have exactly the same form. Thus,
taking into account how the various quantities map onto each other during the
duality transformation, eqs. (4.90) and (4.91) follow from eqs. (4.88) and (4.89).
Since this problem seems so different from the behaviour we find around
the usual connectivity percolation threshold p c , it has been believed that there is no
relation between the critical exponents found in this problem and those of the usual
percolation problem; the two percolation problems are in different universality
classes. This conjecture has been backed by numerical simulations by several groups
[34,35,40,41] Q u r o w n \^e\[ef [s ^ na ^ this problem has the same critical behaviour as
the bond-bending elastic percolation problem. The main argument behind this
conjecture is that there will be effective bond-bending forces appearing in the
network at a larger scale than that of the individual bonds. In fig. 4.17 we show
an example of a configuration where this happens. Our results for the exponents
/ and g, defined through eqs. (4.88) and (4.89) are / = 4.0 ± 0 . 4 ^ and g =
1.29 ± 0 . 0 3 ^ in two dimensions when assuming v = 4/3. These values should be
compared to t + 2v = 3.967 ± 0.005 and s = 1.300 ± 0.005 for the corresponding
critical exponents from the connectivity percolation critical point. The reason for
the discrepancy of our values with those presented by other groups, is that the
numerical determination of / and g is very dependent on the precise value of pv.
The force distribution within the rigid backbone has a multifractal struc-
ture at the rigidity threshold p r , in the same way as the current distribution is
multifractal at pcS42^ Suppose we work in the constant-force ensemble - that is to
say, on each network that we average over in order to determine the statistical
force distribution, we apply a unit force. Other ensembles are of course also possi-
5. Elasticity of central-force systems 153

v
Figure 4.18 A configuration giving rise to an effective bond-bending term. The arrow points at
the bond causing this effect when forces marked F are applied.

ble, such as the constant-displacement ensemble. We compare our results to those


obtained in the electrical conduction problem of section 4.3, and the corresponding
ensemble in that case is the constant-current one.
Let us call the forces appearing inside the network φ. The fundamental
quantity to measure in order to look for multifractality are the moments of the
force distribution,
(<l>k)F~LVF(k). (4.94)
As in the electrical case, the scaling of the zeroth moment gives the fractal di-
mension of the rigid backbone, JDR, and the second moment is the elastic modulus
of the network, so that 2/F(2) = f/u. The fractal dimension was determined to
bet35] DR = 1.64 ± 0.03, which should be compared to DB = 1.62 ± 0.02 for the
usual percolation problemJ13^ The comparison between the second moments we
have already discussed; / « T well within the error bars.
The conjecture that T = t + 2v, eq. (4.84), translates into y^(2) = t/v + 2 =
yi(2) + 2 for the central-force network. It may be generalized to^43^
yF(k) = yi(k) + k. (4.95)
There is no corresponding inequality for these scaling exponents when k φ 2, as
there is for k = 2, see eq. (4.84). However, with the assumption that the central-
force percolation problem is in the same universality class as ordinary connectivity
percolation, eq. (4.94) is the simplest assumption one can make for the connection
between the central-force and connectivity exponents. The conjecture has been
tested up to the fifth m o m e n t ^ and works to within the estimated error bars.
Before we leave the central-force problem, let us mention another contro-
versial subject: The existence of a splay-rigid phase distinct from the rigid phase.
154 A. Hansen / Disorder

It has been argued by Wang et al ^ that there is a second critical point ps be-
tween the connectivity percolation threshold p c , and the rigidity threshold p r , that
distinguishes between a phase with zero Young and shear elastic moduli, but with
a non-zero "Frank elastic constant." In order to give an idea of what splay rigidity
is, think of an industrial design ruler. It may be translated freely along the draw-
ing surface, but not rotated. There has been two numerical studies looking for this
phase, to our knowledge - one with a negative result,^45' and one with a positive
result ^46l In the anisotropic central-force problem,^ Wang et αΐ^ have shown
that such a distinct phase indeed does exist.

4.6 Percolation induced by wide probability distributions


The reader may at this point object that in order to feel the effects of
the percolation critical point in a real physical situation, we would have to be
very close - something that would seem very unlikely to happen in real materials.
However, there are several cases where a system is naturally driven toward the
percolation critical point, and in these cases the percolation theory provides a
powerful framework to understand the phenomenon.
In this section we will describe two such cases where percolation theory is
encountered without fine tuning of parameters. The first one is that of a composite
whose grains have a logarithmically wide distribution of conductances. The second
one is the rupturing of a brittle composite where the local failure strength also
follows a logarithmically wide distribution.
Ambegaokar et αΐ^ demonstrated that if a composite material has a dis-
tribution of local conductances that it is so wide that it is on the verge of not
being normalizable, the conductance of the composite will be that of a percolating
network at the percolation threshold. One such distribution is

Ρα(σ) = (1-α)σ-α, σ€[0,1], (4.96)

in the limit a —> 1. Suppose now that we build up the composite material by first
placing the highest local conductances, then the second highest and so forth. If
we stop this process at any stage and measure the conductance at this point, we
will find that it is smaller than the conductance of the "complete" material. As
a —> 1 in the distribution in eq. (4.96), the ratio between the local conductances at
any stage of the process and those at the next stage, diverge. Thus, in this limit,
the change in the overall conductance becomes more and more insensitive to the
filling-up process - but with one exception: Before we have filled in so many local
conductances that no percolating cluster of conductances exist, the conductance
of the network is zero. At the moment when a percolating cluster is formed, the
conductance makes a jump to the value it has when the material is complete.
Thus, we are dealing with a percolation phenomenon, and the local conductance
6. Percolation induced by wide probability distributions 155

that contribute to the overall conductance of the material is given by the equation
1

Jp1(a)da = pc, (4.97)

where pc is the percolation threshold.


The second, and last example we will quote is that of rupture in a brittle
material, where the local strength of the material has a distribution which is so
wide that it is not normalizable, say, as the one quoted in eq. (4.96) in the limit
a -> lJ48J
Suppose that our brittle material is a composite. We start, say, shearing it.
A stress field builds up inside the material, and as the local grain boundaries reach
their failure stress, they rupture. In a material with a bounded distribution of failure
stresses the question of when a particular grain boundary fail is determined by a
competition between the distribution of failure stresses and the local distribution
of stresses in the material: As the grain boundaries fail, the local stresses rearrange
themselves with the result that grain boundaries with a higher failure fail first.
However, if the failure stress distribution is unbounded, there will always
be grain boundaries that are so weak that they fail no matter what the local stress
field is, as long as there is a stress.
Let us now forego chapter 7, and think of the material as a network where
each elastic bond has a maximum stress it may sustain before it breaks. We start
with all bonds intact and apply a force on the network. If the failure strength is
unbounded, it is therefore enough that a bond belongs to the backbone carrying a
stress in order to be a candidate for breaking.
This breakdown process is therefore closely related to a random dilution
process, which by definition brings the network to the percolation critical point at
the final rupture. The only difference between that process and the present one,
is that the bonds that are picked in that case must be part of the backbone. In
order to investigate how this constraint affects the dilution - or rupture - process,
imagine two identical networks. In one, A, we break bonds at random. In the other
one, B, we pick bonds in the same order as in A, but break only those that belong to
the backbone. In A there is a density of p present bonds - this being the percolation
control parameter. In B the density of present bonds is r. For each r one has a p at
which the backbones of the two networks are identical, and its density is given by
^backbone· I n eq. (4.46), we introduced the fractal dimension of the backbone, DB.
Suppose we have a correlation length £ ~ | p — pc \~u, eq. (4.15), then the density
of the backbone may be related to p — pc as follows

^backbone ~ ^ ~ (p ' Pc)^'^- (4.98)

The rupture process - also called here a screened dilution process - in network B
156 A. Hansen / Disorder

may then be related to the random dilution process in A by

^ = ^backbone ~ (p ~ Vc)"^™'· (4.99)

We now integrate this equation, and find

(r - r c ) ~ (p - i? c ) 1 + I / ( d - D B ) ) . (4.100)

This equation tells us that the screened dilution process is nothing but a percola-
tion process, but with a rescaled control parameter. Inverting eq. (4.100), we may
substitute (p — pc) by (r — r c ) 1 ^ 1 + ^ d _ £ > B ^ in all the power laws we have found near
the percolation threshold pc. For example, if our network is an electrical network,
and the bonds act as fuses, the conductivity of the network will behave as

27 ~ (r - rc)'/(i+^-ßB)e (4.101)

In general, all the scaling exponents related to transport phenomena relating quan-
tities with singular behaviour near p c , to (p — p c ), will be related to the exponents
in the rupture process by dividing them by the factor 1 4- v(d — DB). We have now

come to the end of our survey of percolation theory. If it has been successful, the
reader should have been left with an appreciation of percolation theory as being
a powerful tool in analyzing situations where the disorder is dominant, i.e. the
opposite case of what the homogenization procedures handles well.
The reader should also see the power in the concept of universality - namely
that in the limit of strong disorder, so that the system is close to a critical point,
widely different systems have quantitatively the same behaviour.
The third point the reader should have noted, is the importance of the non-
trivial scaling laws described by critical exponents with values far different from
what should have been expected from simple dimensional analysis.

References
1. Z. Hashin, J. Appl. Mech. 50, 481 (1983)
2. S. Kirkpatrick, Rev. Mod. Phys. 45, 574 (1973)
3. B.I. Halperin, S. Feng and P.N. Sen, Phys. Rev. Lett. 54, 2391 (1985)
4. D. Sornette, J. Physique 49, 1365 (1988)
5. D. Stauffer, Introduction to Percolation Theory (Taylor and Francis, London,
1985)
6. S.R. Broadbent and J.M. Hammersley, Proc. Camb. Philos. Soc. 53, 629 (1957)
7. Phase Transitions and Critical Phenomena, Vols. 1-12, C. Domb and M.S.
Green; C. Domb and J.L. Lebowitz, eds. (Academic Press, London, 1972-1986)
8. J. Feder, Fractals (Plenum Press, New York, 1988)
9. J.M. Normand, H.J. Herrmann and M. Hajjar, J. Stat. Phys. 52, 441 (1988)
References 157

10. J.P. Straley, Phys. Rev. B, 15, 5733 (1977)


11. G.G. Batrouni, A. Hansen and M. Nelkin, J. Physique, 48, 771 (1987)
12. G.G. Batrouni, A. Hansen and S. Roux, Phys. Rev. A, 38, 3820 (1988)
13. H.J. Herrmann and H.E. Stanley, Phys. Rev. Lett. 53, 1121 (1984)
14. A. Coniglio, J. Phys. A 15, 3829 (1982)
15. A. Skal and B.I. Shklovskii, Sov. Phys. Sem. Cond. 8, 1029 (1974)
16. P.-G. de Gennes, J. Physique. Lett. 37, LI (1976)
17. H.E. Stanley, J. Phys. A 10, L211 (1977)
18. H.E. Stanley in On Growth and Form, H.E. Stanley and N. Ostrowsky eds.
(Martinus Nijhoff Publ., Dordrecht, 1986)
19. R. Rammal, C. Tannous, P. Breton and A. -M. S. Tremblay, Phys. Rev. Lett.
54, 1718 (1985)
20. L. de Arcangelis, S. Redner and A. Coniglio, Phys. Rev. B 31, 4725 (1985)
21. T.C. Halsey, M.H. Jensen, L.P. Kadanoff, I. Procaccia and B.I. Shraiman, Phys.
Rev. A 33, 1111 (1986)
22. R. Blumenfeld, Y. Meir, A. Aharony, and A.B. Harris, Phys. Rev. B 35, 3524
(1988)
23. L. de Arcangelis and A. Coniglio, J. Stat. Phys. 48, 935 (1987)
24. S. Roux and A. Hansen, Europhys. Lett. 8, 729 (1989)
25. A.R. Day, R.R. Tremblay and A.-M.S. Tremblay, Phys. Rev. Lett. 56, 2501
(1986)
26. Y. Kantor and I. Webman, Phys. Rev. Lett. 52, 1891 (1984)
27. M. Born and K. Huang, Dynamical Theory of Crystal Lattices (Oxford Uni-
versity Press, New York, 1954)
28. S. Roux and E. Guyon, J. Physique Lett. 46, L999 (1985)
29. S. Roux and A. Hansen, J. Phys. A 21 L941 (1988)
30. S. Roux, J. Phys. A 19, L351 (1986)
31. J.G. Zabolitzky, D.J. Bergman and D. StaufTer, J. Stat. Phys. 44, 211 (1986)
32. H.J. Herrmann and H.E. Stanley, J. Phys. A 21 L829 (1988)
33. Y. Kantor, J. Phys. A 19, L351 (1986)
34. S. Feng and P. N. Sen, Phys. Rev. Lett. 52, 216 (1984)
35. E. Guyon, S. Roux, A. Hansen, D. Bideau, J.-P. Troadec and H. Crapo, to
appear in Rep. Prog. Phys., (1990)
36. A. Hansen and S. Roux, Phys. Rev. B 40, 749 (1989)
37. S. Roux and A. Hansen, J. Phys. A 20, L879 (1987)
38. J. Wang and A. Brooks Harris, Europhys. Lett. 6, 157 (1988)
39. M.A. Lemieux, P. Breton and A.-M.S. Tremblay, J. Physique Lett. 46, LI
(1985)
40. R. Garcia-Molina, F. Guinea and E. Louis, Phys. Rev. Lett. 60, 124 (1988)
41. S. Arbabi and M. Sahimi, J. Phys. A 21, L863 (1988)
42. S. Roux and A. Hansen, Europhys. Lett. 6, 301 (1988)
43. A. Hansen and S. Roux, J. Stat. Phys. 53, 759 (1988)
44. J. Wang and A. Brooks Harris, Phys. Rev. Lett. 55, 2459 (1985)
158 A. Hansen / Disorder

45. R.R. Tremblay, A.R. Day and A.-M.S. Tremblay, Phys. Rev. Lett. 56, 1425
(1986)
46. S. Roux, and A. Hansen, Phys. Rev. B 38, 5170 (1988)
47. V. Ambegaokar, B.I. Halperin and J.S. Langer, Phys. Rev. B 4, 2612 (1971)
48. S. Roux, A. Hansen, H.J. Herrmann and E. Guyon, J. Stat. Phys. 52, 237
(1988)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
©Elsevier Science Publishers B.V. (North-Holland), 1990 159

5. Modelization of fracture in disordered systems

Hans J. Herrmann* and Stephane Roux*

5.1 Introduction
The main goal of this book is the description of fracture of materials in the
cases where disorder is important. One possible approach is to set up some micro-
scopic rules which determine the mechanical behaviour of the medium as well as
how rupture will take place. If the medium is homogeneous, and the rupture deter-
ministic, we can, in general very simply, know what the macroscopic behaviour will
be. However, most systems are more complex, in the sense that they are not ho-
mogeneous and their heterogeneity strongly influences their mechanical behaviour.
How can we face this situation and try to obtain some information at the
macroscopic level starting from a microscopic one and taking into account local
fluctuations due to disorder? This question is very general and appears in many
branches of physics or mechanics. In elasticity of heterogeneous structures, homog-
enization of periodic structures has proved to be very useful. However, no disorder
is present in the description, and since periodicity is a fundamental ingredient of
this technique, it seems hopeless to extract any relevant information concerning
heterogeneities using this tool.
Among the various ways of handling disorder directly, we can, in general,
distinguish three different approaches: mean field, local models, and renormaliza-
tion.
• Mean Field: the mean-field approach consists in neglecting the detailed spatial
arrangement around a given volume element. The surrounding of a volume element
is represented by a homogeneous embedding medium. In this way, each element
interacts equally with all others, and these interactions are represented by the
fictituous "equivalent" homogeneous medium. These conditions imply that mean
field will be a good approximation (1) when interactions are long-ranged, (2) for
problems which are not "strongly" sensitive to the presence of disorder, (3) at high
t Service de Physique Theorique, CEN Saclay, F-91191 Gif-sur-Yvette Cedex, France.
* LPMMH, U.R.A. CNRS 857, Ecole Superieure de Physique et Chimie Industrielles de Paris,
F-75231 Paris Cedex 05, France.
160 H.J. Herrmann and S. Roux / Modelization of fracture

dimensions. For instance, the self-consistent method for determining the equivalent
elastic properties of a heterogeneous elastic composite is a typical example for a
mean-field treatment J1' It gives very good predictions when the disorder is weak,
i.e. when the ratio of the softer to the stifTer component is close to unity. For the
case of random dilution, it still gives reasonable predictions when the concentration
of voids is small. Close to the percolation threshold, the mean-field behaviour is
only valid above the upper critical dimension, i.e. d > 6 (see chapter 4). We will
see that fracture is much more sensitive to disorder than elasticity and therefore,
mean field is expected to give rather poor results in usual space dimensions (two
or three). In addition, fracture naturally enhances the effect of the preexisting
heterogeneities in such a way that the initial disorder cannot be neglected. The
final rupture results from the complex interplay between the initial disorder and
the heterogeneities created by the fracture itself. Some models tend to indicate
that the final stage of rupture can be interpreted as a critical point. ^ This fact
raises serious doubts about conclusions reached about the final stage of rupture
from a mean-field approach.
• Local: a very different spirit stands behind the local approach where one tries to
reproduce as closely as possible the real microscopic behaviour starting from first
principles. This is for instance the case for molecular dynamics. Once the inter-
action between atoms has been chosen, the general molecular dynamic framework
should provide naturally elasticity, plasticity, viscosity, damage, and fracture prop-
erties without having to insert into the model any rheological behaviour by hand.
We will see that this approach applied to fracture is in general so heavy that only
results on small system sizes can be obtained, and therefore no firm conclusions
can be drawn from these studies.
• Renormalization: statistical physics has introduced a very powerful tool to study
critical phenomena which exploits self-similarity: renormalization. It consists in
expressing the invariance of some physical quantities under a change of length
scale. In the framework of field theory, this method has provided many results
for critical phenomena. Although this framework seems rather well fitted to our
problem, attempts in this direction are very difficult in realistic cases. The few
situations where this technique has been used contain some severe approximations,
or use rather unphysical and simplistic geometries.
This pessimistic picture seems to close all doors towards a solution. The
situation is however not so hopeless. Many approaches developed in more details
in the following chapters follow an intermediate route. The spirit of these studies is
to describe very simply the mechanical behaviour of microscopic - or mesoscopic -
elements, and to consider the collective behaviour of these elements in the presence
of disorder (i.e. random variations of the local properties from one element to the
next, or randomness in the initial state or during the time evolution, etc.). This
approach is "local" in the previous sense but it contains some strong simplifications
of the elementary behaviour. We will come back at length on the hypothesis made
in these studies.
2. Classical approaches 161

5.2 Classical approaches


5.2.1 Rupture criteria
The first question to ask when dealing with rupture is to know when a
fracture will develop, and when it will not. A very important criterion has been
proposed by Griffith^ in 1920. The key idea is to formulate an energy balance.
Let us consider a crack of area A in a medium subjected to external forces F
imposed on its boundary dV. Let us now imagine an infinitesimal propagation of
the crack by an amount dA. In this evolution, the potential elastic energy of the
system increases by an amount d£i due to the decrease of stiffness. The work of
the externally applied forces be dW. At the same time, the creation of the new
interface 2dA will cost some energy, d£ 2 = 2^ysdA, where 7S is the surface energy.
The idea of Griffith is to say that the crack will progress, if and and only if the
release of elastic energy is larger than the energy needed to create the new surface
dW > d£i + d£ 2 . Let us call G = dW — d£i/dA the energy release rate. The
Griffith criterion takes the final form: the fracture will progress if the following
condition is fulfilled
G > 27s. (5.1)
The equality in eq. (5.1) defines the critical energy release rate, G c , which controls
the rupture and which is an intrinsic characteristic of the material.
G can be expressed in terms of the forces F and displacements U on the
boundary of the solid:

dV

and

dA
"Wgldi
2 J \ ÖA dA
(5.3,
av '
Let us consider a simple example. In the case of a circular hole of radius a
inside an infinite elastic material of Young modulus E subjected to a traction σ at
infinity, the expression of the elastic energy can be computed exactly, but simple
dimensional analysis already gives the correct answer up to a prefactor:

Sx oc ^f (5.4)

thus using this expression together with £2 oc a 2 7 s , leads to a rupture stress ac:

aQ oc (5.5)

if the imposed stress does not vary during rupture.


Another approach to obtain a criterion for fracture is to consider the stress
in the vicinity of the crack. For a semi-infinite crack in an infinite medium with a
162 H.J. Herrmann and S. Roux / Modelization of fracture

simple stress state imposed at infinity, the stress distribution can easily be obtained
within the framework of linear elasticity. The stress is singular at the tip of the
crack. If we call r the distance to the crack tip, and Θ the polar angle from the axis
of the crack (see fig. (1.9)), the stress varies, close to the crack tip, as
σίά(ν,θ)(χΚφν(θ)/νϊ, (5.6)
where φ^ is a function which depends on the stress state at infinity (mode I, II
or III). The divergence of the stress when the distance goes to zero is very general
and does not depend on the precise geometry. This divergence is unphysical, and
in most cases, elasticity does not describe accurately the state of the material in
the immediate vicinity of the crack tip. We will come back to this point later. Let
us accept this result for a little while. We can characterize the stress state at the
rupture point by the coefficient K of the singularity appearing in eq. (5.6) (K is
normalized to its value for a unit stress imposed at infinity, in each mode). This
coefficient is called the stress intensity factor. Therefore, we will obtain a critical
stress intensity factor for each mode of rupture: K\c, Kuc and Kmc.
In simple geometries, it is possible to calculate the release of elastic energy
of the medium due to the propagation of the crack and to relate this result to
the stress intensity factor. This computation gives the equivalence between this
approach developped by Irwin ^ and the previous one due to GriffithJ3^ Let us
mention the result for plane stress:
Gc = (1/E) {Kl + Kl + (1 + v)Ktllc) . (5.7)
There is another way of estimating the same quantity in plane elasticity
(mode I and II), which is due to Rice^ and has been extented by BuiJ6^ It consists
in calculating a contour integral around a crack tip. The Rice integral, J, is defined
as
J = / ( wenx - OijUj-^1 j ds, (5.8)
c ^ '
where the crack is in the x direction, we is the density of elastic energy, and n is a
unit vector normal to the contour C. The interesting property of this integral is its
independence of the shape of the contour C provided it encircles the crack tip. In
the case of a single crack, this integral is equal to G. In addition, this formulation
allows for a local computation of G which appeared to be a global quantity in its
definition (eq. (5.1)).
Other fracture criteria have been developed^ but they are very rarely used.
Let us simply mention the criterion of Sih ^ which states that the crack will
propagate in the direction of maximum density of elastic energy, and the one based
on the maximum tensile stress GQQ in polar coordinates at the crack tip. We will
not use these criteria in the following.
We have now defined three equivalent quantities, the energy release rate G,
the stress intensity factor K and the Rice integral J. Let us discuss them in more
details.
2. Classical approaches 163

5.2.2 Discussion

The idea of having an energy balance argument is very appealing. However,


one should take into account all physical processes in which energy is dissipated.
The difference dW — d£ x must include all the dissipation of energy due to plastic
deformation, creation of micro-cracks in the neighbourhood of the main crack tip,
viscous loss, etc. In most cases, these different processes which dissipate energy
are themselves very difficult to quantify precisely. Fracture is never an equilibrium
process, and the energy dissipation can very seldomly be neglected. Let us make a
naive comparison. We all know that paper is unstable since if we lit it, it will burn
to ashes with the oxygen of air. Thus if you compare the energy of the initial and
of the final state of the book you are holding in your hand, you will find that it
should rather be smoke and ashes.
Obviously, when a crack has propagated, we can write down an energy bal-
ance which will show that the Griffith criterion was fulfilled. It is however possible
that the criterion is satisfied and that the fracture does not progress anyway. The
Griffith criterion only provides a lower bound on the rupture stress. Many attempts
have been made to correct the original approach in order to describe the additional
dissipation of energy. Let us mention Orowan^ who proposed to add to the surface
energy a contribution due to the plastic work d£ 2 = (27s + 7 p )dA There is however
no way to measure 7 P independently.
Let us consider now the analysis of the stress distribution at the tip of a
crack. We have seen that the solution of linear elasticity leads to the divergence
of the stress close to the crack tip. This is not realistic. If the material is plastic,
there will be a small plastic zone at the tip, whose size, p p , is related to the yield
stress, σ ρ , by pp oc (Κ/σρ)2. This result has been obtained by Irwin.^ Outside
this plastic zone, the behaviour of the material is elastic and the singular solution
of elasticity holds. This plastic radius acts as a short length cut-off of the elastic
solution. The solution of a single crack propagating in an infinite elastoplastic
medium then allows for the introduction of the above mentioned additional term
in the surface energy, as suggested by OrowanJ9^ This contribution represents the
energy dissipated in the plastic flow in the crack tip.
If the material is not plastic but exhibits damage instead as is for instance
the case for ceramics, a small "process" zone will form at the crack tip where
damage will occur. Here too the size, p&, of this process zone will be given by
ρά oc (K/ac)2, where ac is a damage stress characteristic for the material. This
analysis has been proposed by BureschJ10!
Finally, if the material is neither plastic, nor susceptible to damage, Baren-
blatt^11! has proposed to take into account the cohesive forces at the microscopic
level to obtain a local description. The cohesion force in the material originates from
the atomic and molecular interaction forces. At very short ranges, it increases with
the distance between molecules and at longer ranges it decreases. This non-linear
variation can be taken into account by an effective cut-off of the elastic solution.
164 H.J. Herrmann and S. Roux / Modelization of fracture

In all the above cases, the small scale cut-offs are material constants. It is
thus equivalent to measure the stress intensity factor, or the stress at the cut-off
distance. This equivalence will be used at length in the following, where fracture in
disordered systems will be described by a discrete system (which naturally provides
a short cut-off), and where the rupture criterion will be given by a local stress
threshold.
5.2.3 Extension to many cracks
Up to now we have only considered the case of a single crack in an infinite
domain. When many cracks are present, the situation becomes more complex since
the singular stress fields of all cracks interact with each other. The Griffith criterion
can easily be generalized by considering all possible evolutions, i.e. the simultaneous
growth of all cracks, eventually with branching, as well as the nucleation of new
cracks, and computing their energy release rate. The criterion for fracture will be
that at least one configuration releases more energy than necessary to create the
new interface (plus eventual additional contributions coming from plasticity, heat
dissipation, etc.). This approach can however only be executed through numerical
simulations. Let us point out that the numerical procedure is particularity heavy
in this case since all possible evolutions have to be tested.
For many cracks, the criterion based on the stress intensity factor, gives
a different answer than the Griffith criterion. In addition, for two cracks that are
close to each other the very concept of the stress intensity factor looses its meaning.
5.2.4 The importance of disorder
Disorder can come into play in many ways during a fracture process. Spe-
cific cases will be described in detail in the following. Let us just mention here that
generally, fracture enhances an initially present disorder, through the nucleation of
new cracks, or simply due to the heterogeneity of the stress field that results from
the complex geometrical arrangement of the existing cracks. Even small initially
present disorder can be enormously amplified during fracture. Therefore, fracture
is a collective phenomenon, in which disorder plays a fundamental role.
Let us quote three basic experimental facts that will show that disorder is
an unavoidable concept if one wants to deal with real materials (particularly brittle
ones).
• It is well known that statistical fluctuations in the rupture stress of materials
are very often enormous. Two pieces made of the same material, under the same
conditions, will not break at the same time. These sample to sample fluctuations
are present in most materials, and at all scales.
• Besides the statistical fluctuations, the mean fracture stress depends on the
sample size. In general it decreases. Similarly, if the system size remains constant,
but if the characteristic size of the microstructure changes, the mean fracture stress
will be affected. This has been observed for instance in aluminium, where the
microstructure is controlled by the grain sizes, or for ceramics. These size effects
2. Classical approaches 165

tell us that we cannot neglect the small scale structure of the medium.
• Finally, let us have a look at the stress-strain curve drawn schematically in fig.3.6
in chapter 3. This shape is quite frequently encountered in many materials such as
concrete. Let us suppose that we model this material, by describing it at the local
level with this kind of characteristic, within the framework of damage theory. It has
been shown, by Bazant J12^ that in the decreasing part of the curve, called "strain-
softening", the material develops "instabilities". Damage will concentrate in local
zones. If we treat this model numerically with a finite element code, then we will see
that the localization of damage is very dependent on the mesh size used. The finer
the mesh, the more localized the damage. In addition to this geometrical feature,
also the mechanical response depends very much on the mesh size. A homogeneous
solid having this kind of behaviour at a local scale will not have the same behaviour
at the macroscopic scale. Its behaviour is not scale invariant. It can be shown
mathematically, that this instability results from the change of the tangent slope in
the elementary characteristic from positive to negative ("Hadamard instability").
However, since this kind of stress strain relation is observed at a macroscopic scale
for real materials, a stabilizing factor must exist. We believe that disorder indeed
provides such a stabilization.
From these three experimental facts, we learn that it is not possible to
reduce the behaviour of a material to its average. Fluctuations are fundamental,
not only because of the necessary statistical analysis of failure in many practical
instances, but also because these fluctuations at a local scale provide a necessary
stabilizing mechanism to prevent the localization of damage.
What can we do to take into account these effects? Two main philosophies
can be distinguished here. Either we want to describe the mechanical behaviour, or
we want to understand it. Obviously, we would appreciate to do both at the same
time. But the tools that are best suited to one of these approaches will be either
inefficient or very heavy for the other one.
If we want to describe the damage of a strain softening material, we will
have to put into its description an additional contribution to stabilize the localiza-
tion. These additional terms are called localization limitersJ13^ Several proposals
have been made in the past. The principle is very simple, even if the formalism in
its final form may be quite involved. It consists in stabilizing the short length scale
deformations. One crude way is to fix the mesh size of the finite elements, as a
physical parameter. This is not aesthetically satisfactory. One can also add an en-
ergetic contribution which penalizes large strains over short length scales. This can
be done by taking higher order derivatives of the displacement field as suggested
by AifantisJ14! and Schreyer and C h e n ^ Alternatively, one can use a non-local
theory of damage, t16! following former non-local models of elasticity proposed by
K r ö n e r ^ (mentioned in chapter 3). This theory has been developed by Bazant and
Pijaudier-Cabot,^ and applied to the case of concrete by Mazars and Pijaudier-
CabotJ19^ There are still some other ways: S a n d l e r ^ showed that viscous effects
could act as "natural" localization limiters. In all cases, a length scale ξ is added
166 H.J. Herrmann and S. Roux / Modelization of fracture

in the formulation of the problem. In the energetic approach, the coefficient of a


higher order derivative of the displacement can be combined with elastic constants
to give a length scale £. In the non-local approach, the weighting kernel (which
may be for instance a Gaussian function in the space variables) of the non-local
interaction contains a length scale, etc. In these cases £ plays the same role as the
mesh size in a finite element formulation.
This approach allows to reproduce rather accurately the mechanical be-
haviour of strain softening materials and it gives efficient tools for practical appli-
cations. However, it can neither give any information about the length scale since
this is a free parameter of the model, nor about the statistical fluctuations of the
mechanical properties.
If we want to understand the origin of the length scale £ in the model, and
if we want to have some insight about the typical fluctuations of the fracture char-
acteristics, we need to take into account the disorder explicitly in our description.
This type of approach is discussed in the following sections.
Let us emphasize the complementarity of both approaches. Taking disor-
der into account explicitly means that we have to do a lot of statistics, or alter-
natively to deal with large system sizes to describe accurately a situation. This
is computationally very expensive, and cannot be considered as a general way to
handle engineering problems. However, this approach can yield much information
on size effects and the way they should be modelled using a damage theory. The
length scale appearing in damage theory with localization limiters may eventually
be obtained through these methods. The general statistical distribution of failure
probability can also be obtained in this way. Therefore, even if the philosophy of
the approaches is different, the results obtained should be complementary.

5.3 Molecular Dynamics


The most detailed and realistic simulation of fracture is certainly the atom-
istic one. In the ideal case one knows the electronic structure of the atoms and the
crystal structure and can deduce from there the interatomic forces. Using Newton's
law one can then calculate numerically the motion of each atom and watch how the
crystal deforms under an externally applied force. In this way one could monitor
the formation and propagation of a crack on a very microscopic level.
Unfortunately this problem is too ambitious in view of the lack of informa-
tion about the electronic structure and most importantly considering the existing
limitations of computational capabilities. Only recently there have been attempts
to correlate the results of electronic structure calculations with the mechanical
response^21! and one is still far from having a first principles derivation of the in-
teratomic forces that hold the crystal together. For this reason one is forced to use
heuristic interatomic potentials which in the best of the cases can be matched to
experimental neutron diffraction dispersion dataJ22^
Suppose one has chosen a potential. Then a simulation of the system can
3. Molecular Dynamics 167

be performed rather straightforwardly. One considers a finite number of atoms and


starts by placing them on a regular lattice. On the outer boundary of the system
the external load is applied, usually by assigning to the particles that lie on this
boundary a force of fixed value and direction. To obtain the motion of the system
time is discretized into slices At and one calculates the position that each particle
will have after a time At. This can be done in various ways. In the simplest case one
calculates, using Newton's equation of motion, the velocity for each particle from
the forces that act on it due to the interatomic potentials of the surrounding atoms.
After one knows all the velocities v, all particles are simultaneously displaced by a
vector vAt. Then the step is repeated for the next time interval starting from the
new positions of the particles and so on.
Evidently the finiteness of At introduces some error which accumulates
while iterating the procedure and very fast the result could be totally wrong.
To avoid this to happen one has to choose on one hand a small At but on the
other hand one can use more sophisticated methods, for instance using higher
time derivatives ("hash method") or extrapolating by fitting polynomials through
the points ("predictor-corrector method") which can drastically reduce but not
completely eliminate the accumulation of errors. For this reason one has to make
a very careful error analysis to know for how many time steps one can trust the
results. The art of doing this properly is called "molecular dynamics", a field that
has rapidly evolved in the last two decades. Classical methods have been developed
like the "leap-frog" or the V e r l e t ^ method. The convergence can be sped up by
introducing a small amount of damping. We do not want to enter into details here
because these methods have been described in much detail in the literature and
are beyond the purpose of this chapter.
The first applications of molecular dynamics to fracture were done in the
early seventies to study the forces needed to open a single straight crack in a lattice.
Starting with one-dimensional chains modeling the crack surface^24! there were soon
results available obtained for diamond lattices and using rather complicated, angle-
dependent potentialsJ22J In this way predictions can be made for monocrystals of
very specific materials.
Instead of trying to model in more and more detail specific substances it
is also very interesting to understand generic features of fracture that should be
applicable to a wide class of materials although such an approach will not yield to
precise quantitative predictions for given experimental situations. In this spirit a
good and very commonly used interatomic potential is the Lennard-Jones (6-12)
potential
0 y ( r y ) = € [(d/r y ) 1 2 - 2(d/rijf] , (5.9)

where all energies are measured in units of e and d is the equilibrium distance
between particles. For computational reasons this potential is usually cut off at
some distance that could be 1.6rf which includes next-nearest neighbours and which
is well beyond the distance 1.1 Id where the interatomic force is maximum. In this
168 H.J. Herrmann and S. Roux / Modelization of fracture

simplified model neither thermal fluctuations nor quantum fluctuations are taken
into account. Usually the simulations have even been executed in two dimensions
only. The major advantage of these simplifications is that the numerical algorithm
is much faster than in very realistic models and that one can consequently study
systems with more particles.
Within the framework of this more generic model more complex phenom-
ena, like the generation and propagation of dislocations^ could be simulated and
rather non-local properties like the velocity or the surface energy of the crack
could be measured.^ This kind of simulations also lead to the discovery of a phe-
nomenon called "lattice trapping"^ in which contrary to the predictions of the
classical continuum treatment cracks can have a range of stresses for which they
neither propagate nor heal. Measuring simultaneously the change in potential en-
ergy, kinetic energy and surface energy and the work done during crack growth it
was even possible to check the assumptions made on the energy balance by Griffith
for brittle fracture and, surprisingly, important deviations were observedS28^
All above simulations treated the problem of a single crack in a not partic-
ularly disordered medium and it seemed that for this purpose reasonably accurate
results could be obtained by considering up to 11000 particlesJ27^ If one is, how-
ever, interested in how in a rather disordered material microcracks form, interact
with each other and with the disorder of the medium and finally coalesce to break
the entire system apart, one needs a different setup. The attempt to do so will be
discussed nextJ29^
One way of describing a disordered medium is by dilution and this can be
implemented by placing the particles only on a fraction p of the sites of a regular
lattice at the beginning of the simulation. The other sites of the lattice are kept
empty so that one has a concentration c = 1 — p of vacancies. If one applies an
external stress on the particles at the outer boundary some of the vacancies will
open up and grow into cracks. The smallest stress am[n that one needs to apply in
order to break the system apart naturally decreases with increasing concentration
of vacancies and the interesting question is how it decreases and at which con-
centration cc it will vanish. It has been argued ^ that a m j n goes to zero at the
percolation threshold cc = 1 — pc like am[n ~ (p — pc)b and bounds have been given
for the value of b.
In fig. 5.1 we see the results found in ref. 30 doing molecular dynamics
of systems of particles placed on triangular lattices of various sizes. If there are
no vacancies the system reaches a reasonable equilibrium after 6000 iterations
per particle but already for p « | one needs 36000 iterations due to the weak
connectivity. One sees from fig. 5.1 that the data strongly depend on the size of
the system which means that the systems considered are not large enough and that
due to the increasing number of necessary iterations it is not possible to obtain
reliable data close to cc = | . On the other hand, the data point for the 65 x 65 lattice
(+ in fig. 5.1) took already four hours on a Cyber 205 computer. Consequently it is
not possible with the present means of computation to get a numerical verification
3. Molecular Dynamics 169

2.0
I

'mm •

1.5

1.0 -h A i

>

>
0.5 h i l
i > —
A
:
-

I I I
0.1 0.2 0.3 (M 0.5
C

Figure 5.1 Minimum fracture stress amin versus concentration c of vacancies for lattices of linear
size 21 (·), 31 (Δ) and 65 ( + ) (taken from ref. 30). On the triangular lattice the site percolation
threshold is pc = | .
170 H.J. Herrmann and S. Roux / Modelization of fracture

of the predictions for amin at strong dilutions using molecular dynamics.


Although molecular dynamics is the purest, still feasable, discrete approach
it is strongly limited in its applications due to the large amount of computer time
that it consumes. The method works well when a single crack in a regular setting
of particles is studied but the simultaneous presence of several cracks or of disorder
requires such large systems and equilibration times that the method fails to give
good results with the present computing capabilities.

5.4 Lattice models


5.4.1 Why lattice models are used
A numerically tractable, lattice formulation of fracture, alternative to
molecular dynamics, is given by discretizations of continuum equations. In this
case the medium is reduced to a set of points embedded into a grid. Only local
laws, like the balance of force and momentum, are considered and their implemen-
tation involves for each point only a few neighbours. Mathematically the calculation
of collective properties like the equilibrium displacement field is then reduced to
solving a set of coupled linear equations. In chapters 3 and 4 we have already
encountered such examples, like the beam networks or finite element methods.
Evidently these methods do not pretend to describe nature on an atomistic
level as molecular dynamics but have their validity at much larger length scales
where the medium can be described by continuous vector fields. So one does not
have to bother about realistic interatomic potentials because the elastic or plastic
equations of motion of chapter 3 are good enough. The essential behaviour of
the solid will be reproduced by these models in the same sense as the equations
of motion describe real materials even quantitatively if the (phenomenological)
material constants, like elastic moduli, yield thresholds, relaxation kernels, etc.
are properly set. On the other hand, the breaking of the lattice is not a natural
consequence of the simulation as it is the case in molecular dynamics but has to
be put into the model by hand as an additional rule of the model. This rule can
lean on experimental data or on phenomenological laws, several examples will be
discussed, but the fact that it does not follow from first principles shows that lattice
models are less fundamental than molecular dynamics.
One big advantage of the lattice models is that they allow very naturally
for the introduction of disorder. An even bigger advantage is that they are more
"rigid" than a crystal in molecular dynamics because one has fixed neighbours and
the breaking of a bond is an irreversible yes-no decision. Consequently already small
cracks are sharply defined and it is possible to simulate simultaneously many cracks
within rather moderate lattice sizes. Lattice models are therefore good candidates
to overcome the numerical difficulties that we encountered in molecular dynamics.
In engineering practice the by far most commonly used lattice models are
finite element methods (FEM) which have already been mentioned in chapter 3.
We do not want to discuss FEM here in more detail because many pedagogical
4. Lattice models 171

t e x t b o o k s ^ have been written on the subject. The implementation of FEM is


also usually done by using very elaborate, commercially available computer codes.
These codes are tools and used like black boxes rather than being considered to
have a physical interest on their own. The models that we want to discuss in the
following instead are certainly less sophisticated than FEM and some of them can
even be considered to be just special cases of models situated within the framework
of FEM.
The lattice models that have been formulated and investigated recently by
physicists and which are the subject of this book have the advantage that due to
their simplicity one has direct access and eventual physical interpretation to each
of the steps of the algorithm. As a consequence one can in a rather straightforward
and transparent way modify the rules to include the features that will be of interest
to us, like different types of disorder, various connectivities, damage "memory", etc.
5.4.2 General structure of lattice models
All the models that we want to consider here have a certain common setting
which can be summarized in the following way: the medium is discretized such that
all (spatial) sites are equivalent, i.e. the grid is not made finer in regions of higher
stress as is often done for FEM. Each site has the same number of neighbours (co-
ordination number z) so that one has the topological structure of a regular lattice.
The variables that characterize the medium (the electric field for dielectric break-
down, the displacement vector and an eventual local rotation angle for elasticity,
etc.) are placed on the sites of the lattice. The equations that describe the medium
(Laplace for electric models, Lame for elastic models, etc.) are discretized so that
for each site one has one equation per variable which only involves variables on
the z neighbour sites. In the electric case such a local equation is the Kirchhoff
equation. The continuous equations of the medium are therefore transformed into
a set of vN coupled linear equations, where v is the number of variables per site
and N the number of sites of the lattice. Since only nearest neighbours are involved
in each of the equations a solution of the set of equations only involves the inver-
sion of a sparse matrix. Only the boundary condition of the outer boundary on
which the externally imposed constraint is applied is explicitly implemented. This
is done directly in the equations for the sites on these boundaries by expressing
the constraint in terms of the same variables as the ones that one has on the sites.
The boundary condition on the internal boundary, namely the crack surface (or the
surface of the breakdown pattern in the electric case) are automatically fulfilled by
the response of the bonds that constitute the crack which will be discussed a little
further ahead. This response is quantified by the conductivity in electric media,
the elastic moduli in elastic media, etc., and is therefore part of the equations.
After implementing the boundary conditions the set of equations has a
unique solution. Numerically there are many methods to find this solution and the
method that is chosen for a particular case depends on the requirements one has on
the solution. If one needs very high precision because one is for instance interested
172 H.J. Herrmann and S. Roux / Modelization of fracture

in the local electrical potential or strain distribution a conjugate gradient method


eventually sped up by Fourier acceleration^ seems ideal. If one wants to study
dynamic effects like viscoelasticity one can use Jacobi's method giving a physical
sense to a relaxation step. If one does not need precision but wants to save computer
time simple over relaxation gives a fair compromise, although it can under some
circumstances lead to systematic errors.
As will be mentioned in this book at various occasions the simulation of a
rupture process must be done in an iterative way: the equations must be solved in
order to determine which bond (or bonds) should be broken, but once the bonds
have been broken the (internal) boundary condition and therefore also the solution
of the equations is changed. Consequently the equations must be solved again if
one wants to know which bond to break next and so on. We see that the process
is rather heavy because the set of linear equations must be solved as many times
as bonds are broken (unless several bonds are broken simultaneously). With the
present computational means this can be done about one or two thousand times
for a lattice of roughly 104 sites.
The algorithm performed in one iteration can be decomposed into five steps:
(1) solving the set of equations,
(2) determining the set of all the bonds that are eligible to be broken,
(3) calculating for each bond of this set a certain quantity p which is a function of
the solution of the equation on the sites adjacent to the bond,
(4) choosing, according to a rule that depends on p, out of the set, one bond which
will be "broken",
(5) breaking the bond, i.e. changing its conductivity or elastic modulus in the
equations.
Each of these steps allows for a large variety of options that contain many
possible physical situations. (1) describes the nature of the medium and the exter-
nally applied constraints, (2) the connectivity of the crack, (3) and (4) the breaking
rule and the disorder and (5) distinguishes for instance between a breakdown and
a fuse problem. In the following we will discuss each of the steps in more detail.
5.4.3 Equations and boundary conditions
Two types of models have been studied: scalar ones and vectorial ones.
A scalar model can describe an electrical field in dielectric breakdown, pressure
in viscous fingering, temperature in solidification, an electrochemical potential in
electrodeposition, particle density in diffusion limited aggregation, etc. In all cases
one studies an equation which in its most general form is V(eV)</> = 0. In the
electric case φ is the electric potential and e the material dependent dielectric
constant (or tensor). In principle, e can vary from site to site modelizing spatial
inhomogeneities of the material. Choosing for e a random function^ can therefore
constitute a good way of introducing disorder, but we come to talk about this
point later in more detail. In most cases e is considered to be a constant so that
the equation simplifies to Δφ = 0.
4. Lattice models 173

(a) (b)

Figure 5.2 The two most common settings studied in the electrical case applying an external
voltage V; (a) circular geometry, (b) rectangular lattice with periodic boundary conditions in the
horizontal direction and bus bars on top and bottom at which the potential is applied.

Two types of external boundary conditions are usually used in scalar media,
either a circular geometry (fig. 5.2a) or a uniaxial one (fig. 5.2b). The first case is
typically used for dielectric breakdown^34! while the second case has been applied
to random fuse networksJ35^
Vectorial models have been considered up to now only for elasticity al-
though one could also imagine applying them to electrodynamics, fluids, magne-
tohydrodynamics, etc... So, only equations for the displacement field have been
investigated. However, as opposed to the scalar case there are at least two possible
equations for elasticity, namely the Lame eq. (3.12) and the Cosserat eq. (3.14).
In addition, various rather different discretizations for the equations have been in
use: the central-force m o d e l , ^ the bond-bending model ^ and the beam model J38]
These models have been discussed in some detail in previous chapters. The beam
model is a rather straightforward discretization of the Cosserat equations but the
other two models are also believed to describe Cosserat elasticity on very large
scales. The simplest of the three models is certainly the central force model but it
shows pathological behaviour on certain lattices and has very soft modes partic-
ularly on weakly connected graphs. Sometimes also hybrid models, like the Born
m o d e l , ^ have been considered that mix scalar and vectorial terms but their phys-
ical significance is not very clear.
Vectorial models allow for a larger variety of externally imposed constraints
than scalar models. At least four different types of external boundary condi-
tions have been considered: S h e a r ^ (fig. 5.3a), uniaxial tension or compression
(fig. 5.3b), uniform dilation^ (fig. 5.3c) and surface crackingJ41^ The setting of
uniform dilation can be considered as to mimick a membrane spanned on a drum
while the surface cracking was conceived to describe the drying of a layer of paint
on a surface. In this last case each site of the lattice is connected with a spring
174 H.J. Herrmann and S. Roux / Modelization of fracture

DUO (a) (b) (c)


Figure 5.3 Three common ways of imposing an external displacement on the boundary of an
elastic medium; (a) shear, (b) uniaxial tension, and (c) radial dilation, (a) and (b) are square
lattices with periodic boundary conditions in the horizontal direction and (c) is a triangular
lattice.

to another underlying rigid lattice representing the surface which has a somewhat
larger lattice constant as compared to the original lattice. In this way all the bonds
are equally overstretched without having applied any force on a boundary. In fact,
no external boundary is needed in this case so that periodic boundary conditions
can be considered in all directions. We have only discussed two-dimensional ge-
ometries since no serious three-dimensional work has been performed at present
because of the numerical difficulties involved.
5.4.4 Connectivity
There are basically two philosophies one can pursue studying crack prop-
agation: either one wants to understand how one crack grows or one wants to see
how in a stressed medium cracks appear, coalesce and break the system.
The first approach fits into the framework of cluster growth models. For
dielectric breakdown or diffusion-limited aggregation, the connectivity is a natural
consequence of the physical process. This approach can be cast into the formalism
of a moving boundary condition problem as discussed in chapter 8. In order to
implement it into the general algorithm of lattice models described above for frac-
ture one must impose the connectivity of the growing crack by allowing only bonds
on the surface of the crack to be broken. This constraint which is particularly re-
strictive at the early stages of crack growth is rather artifical in most experimental
situations and for this reason this one-crack approach is generally more academic
than the many-crack approach. Nevertheless there are experimental situations that
produce just one connected crack pattern, like the acid-cracking of fig. 1.1c or the
hydraulic fracturing of Van Damme in section 2.6, so that for them the one-crack
approach is well suited. One can also argue that in most cases the last stages of
fracture are determined by the propagation of just one single crack and that then a
one-crack approach could be justified; last stage cracks are, however, usually rather
fast and straight so that on the one hand, they are numerically difficult to handle
4. Lattice models 175

by lattice models and on the other hand they do not generate very interesting, i.e.
ramified or fractal, structures.
If one wants to study the growth of only one crack with a lattice model
one must determine which are the bonds at the surface of the crack, i.e. the set
of bonds eligible for being broken. The surface of a crack in a lattice is, however,
not uniquely defined and so an arbitrary definition of connectivity must be made
which states which bonds in the vicinity of a given broken bond i would, if broken,
constitute together with i a connected crack. Three definitions of connectivity have
been investigated on the triangular l a t t i c e ^ in which one broken bond has 4, 10
and 18 neighbouring connected bonds respectively (see also fig. 8.12).
In the many-crack approach no restriction on connectivity is made so that
at each time step all not yet broken bonds are in principle eligible for being broken.
This approach corresponds to the majority of experimental situations and is used
in chapters 6, 7 and 9.
5.4.5 The breaking rule and disorder
Most of the microscopic physics of rupture is contained in the breaking
rule. As observed in the previous section, the breaking rule has to be put into
the model by hand reflecting the phenomenological nature of lattice models as
opposed to molecular dynamics. The freedom one has in the choice of the breaking
rule and its parameters allows, however, to carefully explore the impact of a given
microscopic mechanism on the macroscopic behaviour of the system. For the sake
of gaining insight into the collective phenomena of rupture it is useful to consider
particularly simple breaking rules but one can also take an engineering point of
view and try to model the material as realistically as possible. In this book we
rather focus on the first approach using simple rules.
As already stressed before at various occasions the effect of disorder on the
breaking process is very important and for this reason its implementation within
the breaking rule should be discussed in some detail: real materials exhibit the
most diverse types of disorder, ranging from small deviations of the crystalline
order, like vacancies, to the large scale heterogeneities of composites. This disorder
can also be in motion: dislocations can migrate, microcracks can form and heal,
interstitials can diffuse, etc. The fundamental distinction that has to be made for
our theoretical purposes is, if this motion is much slower than the investigated
process (i.e. fracture) or not. In the first case the disorder is called "quenched"
and can be considered time-independent. In the second case, called "annealed",
the interplay of disorder fluctuations and fracture has to be taken into account.
On the mesoscopic length scales that lattice models are supposed to de-
scribe the atomistic details of the disorder can be neglected and one can essentially
reduce the description of disorder to a spatial dependence of the local density, elas-
tic modulus or strength. In the case of quenched disorder this spatial dependence is
fixed once and forever before the rupture process starts. Since we ignore the micro-
scopic details and since we adhere to simplicity the spatial dependence is chosen to
176 H.J. Herrmann and S. Roux / Modelization of fracture

I.F.
A »

V- i ►ν.δ

Figure 5.4 Constitutive law chosen for each bond of a lattice model, either in the electric case
(current I against voltage V) or in the elastic case (force F against displacement δ). The breaking
threshold is at Vc or 6C respectively.

be random according to some probability distribution. This distribution contains,


among others, information about the degreee of disorder so that by considering
various distributions the effects of disorder on rupture can be quantified.
In the context of lattice models the individual bond represents the system
on a mesoscopic scale and is therefore characterized by a constitutive law which
in the simplest case should just be linear up to a breaking threshold value as
depicted in fig. 5.4. The strength of the bond is given by the threshold value and
its conductivity or elastic modulus by the slope in fig. 5.4. The presence of disorder
can now easily be implemented by allowing the bonds of the same lattice to have
different constitutive laws. Strength fluctuations can be represented by choosing the
threshold value for each bond randomly.^'43] Spatial variations in the conductivity
or elastic moduli can be described by randomly choosing for each bond different
slopes in fig. 5.4.[33]
The most prominent case of spatial density fluctuations are porous media
where a given point in space is either massive or empty. This situation can be mod-
elled by randomly taking out a bond of the lattice with a probability q = 1 — p be-
fore the breaking process starts and has been extensively studied for electrical^35,44!
and mechanical^ networks (see also next chapter). Particularly interesting is the
behaviour of these "dilute" systems close to the percolation concentration pc.
While in the case of dilution the probability distribution is binary, i.e. the
random numbers are either one or zero, the strength and conductivity fluctua-
tions are generally chosen to be continuously distributed. Most often a uniform
distribution,
V(x) = 1 with 0 < x < 1, (5.10)
is chosen in this case, but also power-law,

V{x) = (1 - α)χα with 0 < x < 1 and a > -1, (5.11)


4. Lattice models 177

and Weibull distributions,

V(x) = ma? m - 1 e-* m with 0 < x, (5.12)


have been considered^ (see also chapter 7). The constants a and m are parameters
of the distribution which determine how strong the disorder is. Distributions with
a lower cut off #/, i.e. for which V(x) = 0 for x < #/, are also of interest ^43l It is
not easy to distinguish experimentally between the different types of distributions.
If one assumes a Weibull distribution values 2 < m < 10 have been fitted to data
measured on rocks, and values of m reaching 20 or even 30 for ceramics (see section
2.1).
In quenched systems the random variables are fixed at the beginning and
the following rupture process is completely deterministic: To break a bond the
equations are solved and the bond for which the ratio strain over threshold 6C
is largest is chosen to be broken. As opposed to this, annealed disorder requires
the production of random numbers during the entire evolution of fracture and
the process is therefore probabilistic. It is not easy to know how to cast a given
microscopic dynamical effect into an explicit expression for such a probabilistic
breaking rule. If the dominant time-dependence of the disorder stems from thermal
fluctuations it seems reasonable to use a Boltzmann factor. So, one approach that
has been takenf41'47^ is to calculate before each breaking for each bond i its elastic
energy e« and from there pi = e -/3ei , where ß = 1/kßT. The bond to be broken,
z'o, is then chosen with a probability proportional to the value of pio. This means
that a random number X is chosen (uniformly distributed) between 0 and Σ ϋ ι Pi
and then the z0 is determined for which Σ ^ 1 Pi < X < Σ?=ι Pi is fulfilled. Instead
of the energy e; also other quantities have been considered as explained in more
detail in chapter 9.
Since thermal fluctuations are not the only source for annealed disorder and
since a first principles derivation of the probabilistic breaking rule is not possible
at present, another approach, inspired from DLA models, has been taken: the bond
to be broken is chosen with a probability pi proportional to the field gradient to
the power η. In the electrical case this means pi oc ιη where i is the current; in the
mechanical case one can consider the strain ε, so that p{ oc ^J 4 2 ' 4 8 ] The exponent
η is phenomenological and can be adjusted to fit experimental results. In general
η is a "relevant" parameter, i.e. the fractal dimensions depend continuously on η.
5.4.6 Other features of breaking rules
As opposed to electric fuses the elastic bonds, represented by beams or
springs, can break in various modes. If the elastic bond is a rubber band it will
tear apart when stretched but if it is a glass rod it will crack when bent. Try,
on the contrary, to break a rubber band through bending or a glass rod through
stretching! In three dimensions another mode of rupture appears, namely torsion.
Mechanical breaking rules should reflect these various rupture modes. This
can be achieved by defining a quantity p that determines the breaking as a sum of
178 H.J. Herrmann and S. Roux / Modelization of fracture

terms each corresponding to one mode. In the quenched case for instance one can
consider that the beam breaks for which
p / n ' + ™x(|mi|,|m,|) (513)
\tfj tm
is largest.^ In this criterion, which is inspired from von Mises' yielding criterion
(see section 3.2.2) applied to beams, / is the force stretching the bond, m\ and
m2 are the moments acting at the two ends of the bond and tf and tm are two
material-dependent threshold values. The first term in eq. (5.13) describes breaking
due to stretching and the second term takes into account the bending mode. If one
has quenched disorder in the strength of the bonds the threshold values tf and
tm are chosen randomly and one can in fact have for tf a distribution between 0
and 1 while tm may be distributed between 0 and s. In this case s measures the
susceptibility of the material to break through bending as compared to stretching.
In the case of annealed disorder one can use the p of eq. (5.13) as the quantity to
which the breaking probability is proportional.
Damage can also be modelled within the breaking rule by lowering the
breaking threshold of bonds in the vicinity of the crack by an amount proportional
to the strain without breaking them. Bonds damaged in this way are more likely
to break at one of the next iteration steps. Short-lived damage can be described
by considering at iteration step t the q u a n t i t y ^
p' = p(t) + foP(t - 1), (5.14)
where p(t) is the usual quantity one would have used at step t without damage and
/o is a parameter. For p(t) one could for instance take expression (5.13). The second
term in eq. (5.14) simulates the memory effect due to damaging that occurred at
the previous iteration step and / 0 controls how strongly this memory acts. The
quantity p' can be used as before either for quenched disorder by breaking the
bond for which p' is largest or for annealed disorder by breaking a bond with
probability proportional to p'.
Remanent damage occurring for instance in stress corrosion is an accumu-
lation of the damaging that occurred during the entire breaking process. It can
be modelled by putting counters c(t) on the bonds susceptible for damaging, like
for instance the bonds on the surface of the crack. At the beginning all counters
are set to zero. At each iteration step t, after having calculated p(t), one defines
ot(t) = (l — c(t — l))/p(t) and chooses the bond which has the smallest ct(t), namely
^min· In the quenched case this is the bond one will break. Then all the counters
are updated by^50^
c(t) = aniBp(t) + fc(t-l). (5.15)
We see that c(t) sums up the damages from all previous iteration steps where the
memory factor / controls how strongly previous damage remains. If / = 1 the
damage is irreversible. The case / —> 0 corresponds to the criterion of eq. (5.14)
with /o = 1. More about these models will be said in section 8.3.3.
4. Lattice models 179

5.4.7 Breaking a bond


Once the bond to be broken has been chosen its actual removal or replace-
ment has to be implemented. In the case of fracture or fuse networks one can simply
take the bond out, i.e. remove the terms corresponding to this bond from the set of
equations describing the medium. This is equivalent to setting the elastic modulus
or the conductivity of the bond to zero.
If one studies dielectric breakdown (DLA), which in two dimensions is the
dual of a fuse network one will instead set the conductivity of the broken bond to
infinity (superconducting).
It is, however, also possible to consider effects of residual strength by just
reducing the elastic modulus or conductivity of the bond by a certain amount.
In this case the interesting phenomenon of crack arrest has been observed^ as
explained in more detail in section 6.5.

5.4.8 General overview of lattice models


We have seen that there is a large variety of lattice models that can be
defined, many of them describing to a certain degree realistic situations. Is there
a standard model which contains the relevant features? The answer is, yes and
no. If one wants to study just the typical mechanisms of crack dynamics of brittle
fracture the random fuse modelt35'44^ which will be discussed in the next chapter is
probably the simplest and best understood candidate. But if one wants to explain
qualitatively a particular experiment this is not enough because there are several
relevant ingredients that considerably influence real fracture. To understand, for
instance, the experiments performed on clay, exposed in section 2.6, the connec-
tivity of the crack is essential. Or, if one wants to reproduce the crack patterns of
drying paint it is necessary to apply the external strain on each site.
It is consequently the richness of the phenomenology of fracture mechanics
that forces us to consider different lattice models. Let us therefore close this sec-
tion by giving an incomplete list of lattice models that have been studied in two
dimensions classifying them according to four criteria:
(1) nature of the equation (sc = scalar, cf = central force, be = beam model)
(2) boundary condition (cir = circular, ax = uniaxial, sh = shear, sur = surface
cracking)
(3) connectivity (0 = no constraint, 1 = only nearest neighbours) (intermediate
numbers correspond to larger neighbourhood)
(4) disorder (st = strength, co = conductivity or elastic moduli, di = dilution,
th = thermal, pr = probabilistic of DLA type)
We list them in table 5.1. There we also show either the fractal dimension found
either for a single crack or, for an ensemble of cracks, we show the exponent that
determines the dependence of the total number of broken bonds as a function of
the system size L. This column illustrates how much the results can vary for the
different models.
180 H.J. Herrmann and S. Roux / Modelization of fracture

nature b.c. conn. disorder d, ref.


sc cir 1 pr 1.7 [34]
sc ax 1 st > 1.1 [52]
sc ax 0 el 1.6 [33]
sc ax 0 di «1 [53]
sc ax 1 pr 1.65 [48]
cf ax, sh 0.8-1 pr 1.25 [48]
cf cir, sh 0.6-1 pr 1.4-1.7 [40]
cf ax 0 th 1.3 [47]
cf sur 0 th fragm. [41]
sc ax 0 di 2* [35,44]
cf ax 0 di 2* [45]
sc ax 0 st 1.7 [43,46]
cf ax, sh 0 St 1.7 [54]
be ax, sh 0 st 1.7 [49]
Table 5.1 Classification of the most important lattice models that have been studied in two
dimensions. The symbol * in column df refers to an actually predicted form L2/yfiuL.

5.5 Renormalization
5.5.1 Introduction
Some physical systems exhibit a particular symmetry, dilation invariance,
which has been highlighted in the study of fractals. This dilation symmetry, also
called "scale invariance", means that the system appears identical under different
magnifications. Besides trivial examples such as homogeneous media, interesting
occurrences of scale-invariant structures appears naturally at critical points, where
a second-order phase transition takes place. The fundamental property of this
symmetry is the absence of a typical length scale.
In order to see explicitly this property, one should consider the "thermo-
dynamic limit", i.e. an infinite system size. Obviously, for all realistic systems, this
scale invariance is confined to a limited range of length scales. There are a lower
and a upper cut-off where this invariance breaks down.
Renormalization consists in building a mapping between one system onto a
similar one considered at a different length scale. If the system is scale-invariant, the
system is mapped onto itself. Therefore, the scale-invariant state can be identified
with the fixed points of the mapping. The identification of the different fixed points
of the mapping allows one to build the phase diagram of the system. A change of
length scale generally affects the physical properties of the system. These changes,
studied at - or in the neighbourhood of - the fixed points permit one to obtain
information on the universal properties of the system. In particular, one can com-
pute critical exponents at a critical point. As mentioned above, among the fixed
5. Renormalization 181

0 1 2

Figure 5.5 Construction of the hierarchical diamond lattice. We show the first two generations
of the iterative construction.

points of the transformation, some are "trivial" (e.g. a homogeneous state) and
some are not (e.g. a critical point). This distinction between trivial or not is not
easy to express precisely, however, the renormalization procedure works in all cases,
whereas for trivial fixed points, generally, there exist more powerful tools to deal
with the properties of these systems.
Renormalization is a very powerful tool, which has been introduced in the
sixties to analyse the universal properties of systems displaying scale invariance,
and in particular critical phenomena. The pioneering work of Wilson in this field,
has been at the origin of his Nobel price.
A good introduction to this method can be found in the textbooks men-
tioned in refs. [55-58]. We will simply illustrate here this powerful tool on a very
simple example for which an exact renormalization transformation can be con-
structed.

5.5.2 A simple example


We consider a "hierarchical" lattice which is perfectly taylored to the renor-
malization transformation. The construction of the lattice is shown on figure 5.5.
One starts with a simple bond (lattice at the zeroth generation). Then one replaces
this bond by four bonds connected like a diamond (lattice at the first generation).
Then each of the four bonds is replaced by a diamond of four bonds again, and so
on ad infinitum.
This construction defines the hierarchical lattice, on which we will study
percolation, as discussed in the previous chapter. Therefore, we remove at random
a fraction (1 — p) of bonds and study the connectivity properties of the resulting
structure.
As mentioned before, we need to define the renormalization transforma-
tion. In order to do so, we note that the hierarchical construction of the lattice is
well-fitted to the scale transformation, since the lattices map on themselves when
182 HJ. Herrmann and S. Roux / Modelization of fracture

the length scale is changed by an integer power of two. Thus, to establish the cor-
respondence, from one size to another, we
• first map each diamond of four bonds into one macro-bond, and then
• scale down the size of the lattice by two. In this way we will be able to map the
lattice onto itself.
In the case of a random dilution of the lattice, we have now to compute
the probability, p\ for a "macro-bond" to be present. Since we want to preserve
the connectedness property of the original lattice, we choose to have a "macro-
bond" present, if and only if the four bonds it will stand for were such that the
two ends of this elementary diamond were connected. In terms of probability, it is
straightforward to compute p' as a function, Τ(ρ), of p:

F(p) = l-(l-p2)2. (5.16)

The renormalization transformation maps a diluted lattice, with probability p onto


a similar one with probability T{p).
Figure 5.6 shows a schematic graph of T{p). One sees that \ί ρ is close to
one, T{p) will be closer to one, whereas if p is close to zero, Ρ(ρ) will be closer to
zero. Iterating the renormalization transformation, most cases will flow towards p =
1 or p = 0. These two attractive fixed points are necessarily separated by a repulsive
one, p = pc = (\/5 — l ) / 2 . As mentioned above, all these fixed points correspond
to scale-invariant states. We see from this simple example that two of these fixed
points are "trivial": p = 0 and p = 1. They correspond respectively to an empty
and a full lattice. The unstable fixed point, pc, is exactly the percolation threshold
(introduced in chapter 4) of the hierarchical lattice. In this example, the scale-
invariance of this fixed point is to be taken in a statistical sense. To use the language
introduced in the previous chapter, if p is different from the percolation threshold,
the correlation length, £, is finite, and under the renormalization transformation, we
scale it down by a factor of two. Upon iteration of the transformation, we push the
correlation length ξ to zero and thus we have access to the properties of the lattice
at larger and larger scales, and we see the properties of a homogeneous medium
(the trivial fixed points). At threshold the correlation length is infinite, and this
absence of a characteristic length scale is the essence of the scale invariance.
Suppose we are in the neighbourhood of the percolation threshold, p =
pc + e, with e <C 1. Under the renormalization transformation, we will escape
from threshold like p' = pc + e', where e' = edT(p)/dp\Pc. At the same time, the
correlation length, which initially was £, will be scaled down to ξ' = ξ/2. The
transformation of p and ξ close to pc allows to extract their singular relationship:

ί(ρ)~|ρ-ΛΓ" (5-17)

since

l=2=(i) A
"' (518)
5. Renormalization 183

p T
1 l·

Pc

^—-"^ < — ■ — ►
0 p 1 P

Figure 5.6 Schematic plot of the function T{p). The intersection of the curve with the diagonal
gives the three fixed points of the renormalization transformation: p = 0, p c , and 1.

where λ = d!F(p)/dp\Pc. In the case of the diamond lattice, we compute λ =


2(3 - Λ/5) from eq. (5.16) and thus v = log2/logA « 1.635.
Therefore, we have able in this simple case, to show a renormalization
transformation at work. We have obtained the simple "phase diagram" (the fixed
points), and one critical exponent (that of the correlation length). All other critical
properties discussed previously in chapter 4 can be worked out on this simple
geometry. We leave this exercise to the reader.
Let us finally mention that we have chosen an example fitted to our pur-
pose, where a simple, yet exact, renormalization transformation can be found. In
more complex cases, particularity on Euclidian lattices, it is not always possible to
find an exact renormalization transformation. This difficulty is the most restric-
tive limitation of renormalization in practical cases. We will describe next some
attempts to apply renormalization to the case of fracture performed on hierarchical
lattices, in order to gain some insight on the large scale behaviour of disordered
system. It is unfortunately restricted to very simplistic cases, and still, the result
is not "trivial."
5.5.3 Application to fracture
Let us consider the application of scale invariance to rupture of disordered
systems. As mentioned previously, the studies that have been carried out in this
direction are rather simple. The geometries considered are often very unrealistic,
because they are taylored to the method, and the stress redistribution inside the
medium often artificial, or approximative. However, the method illustrates quite
nicely a few important concepts that appear in more realistic cases.
The first pioneering work in the field is due to Smalley et α/J59,60^ Renor-
184 H.J. Herrmann and S. Roux / Modelization of fracture

malization was used to study a problem close to the one initially introduced by
Harlow and PhoenixJ61^ Smalley et α/t59,60^ studied a hierarchical model of fracture
which was designed to describe the propagation of a crack along a geological fault.
The fault consisted initially of an array of asperities, each of them being character-
ized by a failure stress ac. The simplest version of this model is one-dimensional,
although the two-dimensional generalisation of it is straightforward. The failure
stress is randomly distributed according to a distribution /ο(σ € ). The asperities
are grouped in pairs, which are themselves grouped in pairs again, and so on. This
hierarchical structure of successive pairs is fundamental for the following. If a pair
of asperities supports a stress σ then the load is shared equally between the two
elements. If one of these asperities fails, then the other must support the total
load. This procedure is repeated from generation to generation to redistribute the
total load. This type of sharing of the load is rather artificial, since it does not
take into account the variations in asperities over more than one generation. We
can, however, easily write down the probability, Ρ„(σ), that a group of asperities
at generation n can sustain a given load σ as a function of the probability at the
previous generation:

Ρη(σ) = Ρη2_χ(σ) + 2Ρ„_ 1 (2σ) (1 - Ρ„_ α (σ)). (5.19)

This equation contains all the necessary information to proceed further. At the
zeroth generation, Po(x), is nothing but the integral of fo(x). We are interested in
the asymptotic behaviour of Pn{x) for large n. The key to the answer lies in the
fixed point of the transformation (5.19).
Despite the simplicity of the model, the answer to the previous questions
is not straightforward. Smalley et alJ59'60^ initially proposed some approximation
in projecting the transformed distribution at each iteration onto a Weibull type
distribution. They obtained that Ρη(σ) tends to a Dirac function centered around
a well defined value of σ. In addition, the width of the distribution decreases as
a power law of the system size. A similar problem was considered by SornetteJ62^
who obtained different regimes depending on the initial distribution fo(x).
Finally, a mathematically rigorous treatment of this question has been pro-
posed recently by Gabrielov and Newman,^ together with some numerical sim-
ulations: Under some hypothesis concerning the initial distribution Po(x), (i.e. if
P o (0) = 1, Po(x) < 1 for all x > 0, and Po(x) goes to zero faster than 1/x) the
stress ση for which the probability to support the load is Ρη(ση) = 1/2, goes to
zero as
ση oc - oc -— — (5.20)
log n log(log L)
where L = 2 n is the system size. This very weak size dependence differs from the
previous result of Smalley et al., who obtained that ση is finite.
This result is quite strong, and shows one particularity of the fracture
problem. Whatever the mean strength of the elements constituting the structur,
References 185

(here asperities), if there are bonds of arbitrarily small strength (second hypothesis
on the initial distribution) then the mean strength of the system will vanish in
the thermodynamic limit. The weakest bonds, even being very rare, completely
control the global behaviour. The mathematical proof of the above result is rather
complicated. The origin of this difficulty lies in the very nature of the problem. We
have to deal with recursion relations and a fixed point in the space of functions
(here the probability distribution Ρ η (σ)), whereas in usual renormalization group
treatments, one works with scalar quantities (see e.g. the probability p' in the
previous example on the hierarchical lattice). All this despite the fact that the
rules for the load sharing are particularly simple in this example. In more realistic
models, the redistribution of the load due to the rupture of one element does not
simply happen within one level of the tree-like structure.
Other attempts have been made changing the lattice structure, or the load
distribution rules. We have already mentioned the work by S o r n e t t e ^ on a hierar-
chical diamond lattice which leads to a recursion relation very close to eq. (5.19).
Sornette and Redner ^ have also studied the so called "bubble" model which
consists of a long string of L bundles in series, each bundle being made of w fibers
in parallel. Depending on the relative values of L and w, various scaling regime
were obtained.

5.5.4 Conclusion
An exact renormalization method for fracture is quite complicated for the
following reason. In real cases, one has to know the complete characteristic σ — ε
of each constitutive element. Due to the stochastic nature of the problem, these
characteristics appear with a certain probability. We thus have to deal with the
transformation of probability measure on the space of characteristics.
One sees here the difficulty of renormalization. In the first example, we had
to deal with a simple scalar quantity: the probability p on the hierarchical model.
In the latter case, the simplicity of the geometry and of the load sharing rules
allowed to deal only with a probability distribution of a scalar number, i.e. the
failure stress. In the general case, we have to handle probability distributions over
a set of characteristics. Let us note, however, that this kind of problem is implicit
to legitimate the basic assumptions of "damage theory" introduced in chapter 3.
Let us hope that in the future a more complete treatment of this complex case will
be possible.
We acknowledge the support of the ATP "Materiaux Heterogenes" of the
PIRMAT.

References
1. E. Sanchez-Palencia and A. Zaoui eds., Homogenization Techniques for Com-
posite Media, (Springer-Verlag, 1986)
186 H.J. Herrmann and S. Roux / Modelization of fracture

2. L. de Arcangelis, A. Hansen, H.J. Herrmann and S. Roux, Phys. Rev. B 40,


877 (1989)
3. A.A. Griffith, Phil. Trans. Roy. Soc. 221, 163, (1920)
4. G.R. Irwin, Appl. Mat. Res. 3, 65 (1964)
5. J.R. Rice, J. Appl. Mech. 35, (1968); see also J.R. Rice in Fracture, vol. 2, H.
Liebowitz ed., (Acad. Press, 1968)
6. H.D. Bui, Eng. Fract. Mech. 6, 287 (1974)
7. H.D. Bui, Mecanique de la Rupture Fragile, (Masson, Paris, 1978)
8. G.C. Sih and H. Liebowitz, in Fracture, vol. 2, H. Liebowitz ed., (Acad. Press,
1968)
9. E. Orowan, Welding J. Res. 34 Suppl., 1575 (1955)
10. F. Buresch, in Fracture Mechanics of Ceramics, vol. 4, R.C. Bradt, D.P.H.
Hasselman and F.F. Lange eds., (Plenum Press, New York, 1978), p. 835
11. G.I. Barenblatt, Adv. Appl. Mech. 7, 55 (1962)
12. Z.P. Bazant, ASME Appl. Mech. Rev. 39, 675 (1986)
13. Z.P. Bazant and T.B. Belytschko, Int. Conf. on Constitutive Laws for Engi-
neering Materials, C.S. Desai et al eds., (Elsevier, New York, 1987)
14. E.C. Aifantis, J. Eng. Mat. Tech. 106, 326 (1986)
15. H.L. Schreyer and Z. Chen, J. Appl. Mech. 53, 791 (1986)
16. Z.P. Bazant, T.B. Belytschko and T.P. Chang, J. Eng. Mech. Div. ASCE 110,
1666 (1984)
17. E. Kröner, Int. J. Sol. Structures 3, 731 (1967)
18. Z.P. Bazant and G. Pijaudier-Cabot, J. Appl. Mech. 55, 287 (1988)
19. J. Mazars and G. Pijaudier-Cabot, J. Eng. Mech. Div. ASCE, to appear (1990)
20. I.S. Sandier, Proc. Symp. on Constitutive Equations, ASME, K.J. Williams
ed., (New York, 1986)
21. C.L. Briant and R.P. Messmer, Philos. Mag B 42, 569 (1980);
M.E. Eberhart and D.D. Vvdensky, Phys. Rev. Lett. 58, 61 (1987)
22. J.E. Sinclair and B.R. Lawn, Proc. Roy. Soc. London A 329, 83 (1972)
23. L. Verlet, Phys. Rev. 159, 98 (1967)
24. R. Thomson, C. Hsieh and V. Rama, J. Appl. Phys. 42, 3154 (1971)
25. J.H. Weiner and W. Pear, J. Appl. Phys. 46, 2398 (1975)
26. W.T. Ashurst and W.G. Hoover, Phys. Rev. B 14, 1465 (1976)
27. see e.g. D.M. Esterling, Comm. Solid State Phys. 9, 105 (1979); or
A. Paskin, D.K. Som and G.J. Dienes, J.Phys.C 14, L171 (1981)
28. A. Paskin, A. Gohar and G.J. Dienes, Phys. Rev. Lett. 44, 940 (1980)
29. B.K. Chakrabarti, D. Chowdhury and D. Stauffer, Z. Phys.B 62, 343 (1986)
30. P. Ray and B.K. Chakrabarti, J. Phys.C 18, L185 (1985) and Solid State
Comm. 53, 477 (1985);
See also review by B.K. Chakrabarti, Rev. of Sol. State Sei. 2, 559 (1988)
31. S.S. Rao, The Finite Element Method in Engineering, (Pergamon Press, Ox-
ford, 1982) and P. Tong and J.N. Rossettos, Finite Element Method: Basic
Technique and Implementation, (MIT Press, Cambridge, 1977)
References 187

32. G.G. Batrouni and A. Hansen, J. Stat. Phys. 52, 747 (1988)
33. H. Takayasu, in Fractals in Physics, eds. L. Pietronero and E. Tossatti (Elsevier,
Amsterdam, 1986), p.181 and H. Takayasu, Phys. Rev. Lett. 54, 1099 (1985)
34. L. Niemeyer, L. Pietronero and H.J. Wiesmann, Phys. Rev. Lett. 52, 1033
(1984)
35. L. de Arcangelis, S. Redner and H.J. Herrmann, J. Physique Lett. 46, L585
(1985)
36. S. Feng and P.N. Sen, Phys. Rev. Lett. 52, 216 (1984)
37. Y. Kantor and I. Webman, Phys. Rev. Lett. 52, 1891 (1984)
38. S. Roux and E. Guyon, J. Physique Lett. 46, L999 (1985)
39. H. Yan, G. Li and L.M. Sander, Europhys. Lett. 10, 7 (1989)
40. E. Louis, F. Guinea and F. Flores, in Fractals in Physics, eds. L. Pietronero
and E. Tossatti (Elsevier, Amsterdam, 1986) and E. Louis and F. Guinea,
Europhys. Lett. 3, 871 (1987)
41. P. Meakin, Thin Solid Films, 151, 165 (1987)
42. P. Meakin, G. Li, L.M. Sander, E. Louis and F. Guinea, J. Phys. A 22, 1393
(1989)
43. B. Kahng, G.G. Batrouni, S. Redner, L. de Arcangelis and H.J. Herrmann,
Phys. Rev. B 37, 7625 (1988)
44. P.M. Duxbury, P.D. Beale and P.L. Leath, Phys. Rev. Lett. 57, 1052 (1986);
P.M. Duxbury, P.L. Leath and P.D. Beale, Phys. Rev. B 36, 367 (1987)
45. P.D. Beale and D.J. Srolovitz, Phys. Rev. B 37, 5500 (1988)
46. L. de Arcangelis and H.J. Herrmann, Phys. Rev. B 39, 2678 (1989)
47. Y. Termonia, P. Meakin and P. Smith, Macromol. 18, 2246 (1985)
48. E.L. Hinrichsen, A. Hansen and S. Roux, Europhys. Lett. 8,1 (1989)
49. H.J. Herrmann, A. Hansen and S. Roux, Phys. Rev. B 39, 637 (1989)
50. H.J. Herrmann, J. Kertesz and L. de Arcangelis, Europhys. Lett. 10, 147 (1989)
51. Y.S. Li and P.M. Duxbury, Phys. Rev. B 38, 9257 (1988)
52. F. Family, Y.C. Zhang and T. Vicsek, J. Phys. A 19, L733 (1986)
53. P.D. Beale and P.M. Duxbury, Phys. Rev. B 37, 2785 (1988)
54. A. Hansen, S. Roux and H.J. Herrmann, J. Physique 50, 733 (1989)
55. T.W. Burkhardt and J.M. J. Van Leuwen, Real-Space Renormalization, in Top-
ics in Current Physics, vol. 30, (Springer, Berlin, 1982)
56. T. Niemeijer and J.M.J. Van Leuwen, in Phase Transitions and Critical Phe-
nomena, C.Domb and M.S. Green eds., (Acad. Press, London, 1972)
57. B. Widom, J. Chem. Phys. 43, 3892 and 3898 (1965)
58. L.P. Kadanoff, Physics 2, 263 (1966)
59. R.F. Smalley, D.L. Turcotte and S. Solla, J. Geophys. Res. 90, 1894 (1985)
60. D.L. Turcotte, R.F. Smalley and S. Solla, Nature 313, 671 (1985)
61. D.G. Harlow and S.L. Phoenix, Adv. Appl. Probability 14, 68 (1982)
62. D. Sornette, J. Physique 50, 745 (1988)
63. W.I. Newman and A.M. Gabrielov, Failure of hierarchical distributions of über
bundles; I, preprint;
188 H.J. Herrmann and S. Roux / Modelization of fracture

A.M. Gabrielov and W.I. Newman, Failure of hierarchical distributions ofnber


bundles; II, preprint
64. D. Sornette and S. Redner, J. Phys. A 22, L619 (1989)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
©Elsevier Science Publishers B.V. (North-Holland), 1990 189

6. Breakdown of diluted and hierarchical systems

Phillip M. Duxbury*

6.1 Introduction and overview


There has been intense study of relationships between the microstructure
and properties of materials (structure-property relationships). A deeper under-
standing of the effects of disorder on structure/property relationships has de-
veloped over the last two decades by a combination of effective-medium theo-
ries, scaling theories, numerical analysis and well controlled experiments on model
systems J1-3^ Most of this progress has been on transport^1,2^ or elastic^1,3! (low mo-
ment) properties, while breakdown in random media (extreme moment properties)
has been virtually ignored. This book is a first attempt to remedy that omission,
and this chapter in particular is devoted to calculating the breakdown properties of
materials with composite (two component) or hierarchical microgeometries. These
microgeometries occur in a variety of physical and technological materials, and play
a central role in the pedagogical development of the theory of disordered systems
f1-3^. In this chapter we use them to illustrate the effects of disorder in breakdown
phenomena.
We first set up the notation and ideas for the study of breakdown in disor-
dered systems by using single crack results to illustrate that breakdown properties
are much more sensitive to defects than are conductivity, elastic moduli and other
transport properties. This is especially true of brittle materials such as ceramics,
where it is known that a small crack (e.g. 0.50 mm <C sample size) has negligible
effect on the elastic constants but can drastically reduce the tensile strengthJ4^ In
section 6.2 we demonstrate this, and also demonstrate the connection between low
moments and transport properties/elastic moduli and between extreme moments
and brittle strength. We then show that the crossover between these two limit-
ing cases occurs at a critical value of the moment order, rac, that depends on the
sharpness of the crack tip, the crack length and system sizeJ5^
The defect sensitivity of breakdown properties suggests that a small amount
of disorder will have a much stronger effect on (brittle) breakdown strength than
t CFMR and Physics/Astronomy, MSU, East Lansing, MI 48824-1116, USA.
190 P.M. Duxbury / Breakdown

on transport and elastic moduli. In section 6.3 we illustrate this effect by discussing
network models for breakdown in random mediaJ 6-71 l In particular, we concentrate
on networks with random dilution, which we study using a novel extreme scaling
analysis I34'38] in combination with numerical simulations. This analysis leads to
the prediction of a logarithmic size effect and a singular behaviour in the limit
of low porosity. The analysis is based upon the assumption of brittle network
failure caused by the propagation of one dominant crack. Within this dominant-
crack theory, the analysis can be extended to the calculation of fluctuations about
the average strength and indeed the complete distribution of strengthsJ38»40! This
distribution has considerable practical importance due to its use in extrapolating
from test load conditions to design loading conditions. Most often the Weibull
distribution is used for this extrapolationJ 74 ' 75 ! yet a detailed analysis of diluted
networks shows that for these disordered systems, a modified Gumbel distribution
is more appropriateJ 38 ' 40 ! In addition to outlining the derivation of this reliability
distribution for diluted networks, we discuss possible tests to differentiate between
it and the Weibull form.
If the failure of a disordered material is not dominated by a single crack
of size a <C L (= system size), a different analysis is necessary. One such class
of problem is materials and models with a fractal or hierarchical geometry. In
these problems, a scaling or renormalization-group analysis similar in spirit to that
used in the analysis of critical phenomena near second order phase transitions is
applicable (see also section 5.5). We illustrate this technique with a scaling analysis
near the percolation pointt34'45-47,56^ in diluted networks and with calculations on
hierarchical systemsJ 24,33,69-71 ^
Another problem not amenable to the "dominant crack analysis" is a disor-
dered system where a great deal of local damage occurs before macroscopic failure.
One important aspect of this class of problem, is that the greater the damage
before failure, the greater is the energy absorbed and hence the greater the tough-
ness. For these problems, no comprehensive analytic analysis is yet available. We
thus illustrate the open issues with numerical r e s u l t s ^ and with a few analytic
results on rather special solvable systemsJ60'72! Some of the network models with
continuous distributions of bond strengths^ 32 ' 60-63 ] to be discussed in chapter 7 of
this book are dominated by a damage mechanism rather than brittle failure and
again have not yielded to a complete analytic analysis.
In section 6.6, we discuss four experimental issues to which the theoretical
modelling has direct application. The chapter closes in section 6.7 with a brief
summary and overview.

6.2 Scaling theory for an isolated crack


6.2.1 Moduli and breakdown strength
From a continuum analysis^ it is known that an isolated crack of size
a <C L (L is the linear dimension of sample), has the following effect on Young's
2. Scaling theory for an isolated crack 191

modulus, £?, and electrical conductance, Σ:

EeS(a)/E0 ~ 1 - 0(1/Ld) (6.1)

and
Σβζ(α)/Σ0 ~ 1 - 0(1/Ld). (6.2)
Here subscripts "eff" refer to the effective properties of a system containing a
crack, while subscripts u 0" refer to systems without a crack. In contrast,t5,76^ a
stress intensity analysis leads to the following formulae

ac(a)/ac(0) ~ («/α) 1 / 2 (6.3)

for the tensile strength, σ0, of brittle material undergoing mode I failure (see chap-
ter 1, where the various deformation modes are discussed), a is the crack length and
K is the crack tip curvature. Similarly if a conductor fails when a current greater
than a critical value, j c , is applied across it (fuse behaviour) then

ic(a)/j c (0) « (κ/α) 1 / 2 (6.4)

(if a local Joule heat criterion is used the exponent in equation (6.4) changes to
a 1).
On comparing equations (6.1) and (6.2) with equations (6.3) and (6.4) it
is seen that breakdown properties are far more stongly affected by a large crack
(with a <C L) than are transport or elastic moduli.
To illustrate how finite system size affects this conclusion, we present in
figure 6.1 the effect of a crack of size a on electrical strength and conductivity in a
system of size L.
For a/L <C 1 these results are consistent with equations (6.1) - (6.4), while
for a/L larger they include the boundary effects. Again the sensitivity of brittle
strength to small flaws is clearly evident.
6.2.2 A moment spectrum analysis of crossover effects
The difference between the scaling behaviour characteristic of moduli (equa-
tions (6.1) and (6.2)) and that of brittle strength (equations (6.3) and (6.4)) is due
to the greater sensitivity of strength to load hotspots. Consider then a distribution
function that gives the probability, P(x), that a load, #, occurs in the material.
Since brittle strength is determined by the largest values of x, x m a x , we have (for
the mechanical case x = σ),

ac(a)/ac(0) - a0/amaLX(a) - J m ^ o / ((σ™)1'™) , (6.5)

where as usual
(am) = J P(a)am da. (6.6)
192 P.M. Duxbury / Breakdown

a)

1.0 ' »+».Η


^ ~ —

0.8 k

0.6

+ J
0.4 +
+

'♦♦.
<><>,ooo
0.2 l· oo°oo<>
°ooo ^Oo,
o«o,oo,
OoÄ
0.0
0 0.2 0.4 0.6 0.8

Figure 6.1 a) The geometry of a conductor containing a crack and; b) the conductivity (+) and
breakdown current (0) as a function of the length, a, of a (lattice) crack in a network of size L.
The calculations are for a crack of width one lattice spacing in 80 x 80 square networks, using a
local current threshold criterion for the irreversible failure of a bond.
2. Scaling theory for an isolated crack 193

Brittle strength is clearly related to the very high moments of P(x), and for this
reason is called an extreme moment quantity. In contrast, we have

total energy of network = LdE0 IΡ(σ)σ2 άσ = LdEeff(a)al, (6.7)

then,
Eeff(a)/E0 = σ20/(σ2) (6.8)
so that Eeff is related to the second moment of the distribution Ρ(σ). Similar
arguments apply to Zeg and jc(a). By studying the form

Rm = x0/(xm)1/m (6-9)
(where x = σ or x = I) as a function of m, we should see a crossover from low
moment scaling (such as equations (6.1) and (6.2)) to extreme moment scaling
(such as (6.3) and (6.4)). As well as being of fundamental interest, this critical
moment value will determine the sort of measurements that are likely to provide
good indicators of the presence of subcritical flaws (a subcritical flaw is one that
is slowly growing, but which is not sufficiently large to cause catastrophic failure).
To determine the scaling behaviour of Rm as a function of moment order,
we must evaluate (xm) in a system containing a single crack. This is not exactly
calculable, but may be estimated from the following asymptotic forms for the field
near a crack-like void in uni-axial tension^
[ 1 + a2m/rd if r > a;
σζ(ν)/σ0 ~ < clm + c 2 m (a/2r) 1 / 2 if κ < r < a; (6.10)
1 hm + k2rn(al2n)112 if r < K.
Here the stress is applied along the z direction in a 2-d system containing a crack
with its long axis in the x direction, r is the distance from the crack tip along the
x direction. a 2 m is a dipole moment for the 2-d problem of a crack in an elastic
background, κ is the crack tip curvature (= b2/a) for the crack-like ellipse (of semi-
axes a and b) and Cim, C2m, k\m and k2m are constants. The scaling behaviour of
equation (6.10) is quite universal, and applies (with different a, c and k) to ellipse,
ellipsoidal and slit-like cracks in 2-d and 3-d linear elastic and electric problemsJ5^
As the moment order m increases the regions r < a in equation (6.10)
increase their contribution to the integral (6.6) and as m —► oo the hotspots dom-
inate. This is explicitly seen by subdividing and approximating the integral (6.6)
as,

V([a(r)/a0]m) * j rd~l dr(l + a2Jrd)m


r>a

+ J r d r [ c l m + c 2 m (a/2r) 1 / 2 ] m
K<r<a

+ j r dr[klm + fc2m(a/2«)1/2]m. (6.11)


r<K
194 P.M. Duxbury / Breakdown

where V = Ld is the volume. Keeping the most singular terms in each integral, we
find;
(6.12)
The mO(l/V) term is the leading order term in the expansion of the dipole integral
in equation (6.11) (to do this properly the angular terms must be included). From
equation (6.12), we can see that for fixed a and κ, Rm has the following scaling
forms
(6.13)

where crossover from low moment type behaviour to extreme moment behaviour
occurs at rac given by
(6.14)
or
(6.15)
Due to the universality of the forms given for the mechanical case in equation (6.10),
this result straightforwardly generalizes to three-dimensional elastic problems and
to electric problems in two and three dimensions.
There is also a great deal of conceptual similarity between the problem of
an insulating crack in a conducting background and the problem of a conduct-
ing crack in an insulating background^5,70! (see also chapter 3 of this book). The
former problem has been discussed as the simplest example of an electrical break-
down problem f28'38'40], while the latter is the simplest starting point for estimating
the dielectric breakdown field as a function of the volume fraction of conducting
inclusionsJ38,40,70! It is well known that (e.g. by duality - see chapter 3) in two
dimensions, the electric field enhancement at the tip of a metal ellipse oriented
parallel to the applied electric field £ 0 m an insulating background, scales in the
same way as that at the tip of a void inclusion oriented perpendicular to the applied
field in a conducting backgoundJ5! The asymptotic form (6.10) then applies equally
well (with the replacement of stresses by electric fields) to the dielectric problem in
two dimensions, and hence equations (6.12) to (6.15) express the moment scaling
appropriate to that problem. The dielectric problem is different in three dimensions
however, as there the most important crack-like defect is a finger-like inclusion f5,34^,
which induces the following electric field, £, behaviour;^

(6.16)

In this case the asymptotic evaluation of the integral (6.11) gives


3. Extreme scaling analysis of dilute networks 195

+ / r 2 dr[ln(r/2a) + a/2r] m
K<r<a
+ « 3 [l + (a/6) 2 /ln(2a/6)] m (6.17)

so that
([£(r)/£ 0 ] m > ~ 1 + mO(l/V) + (a/K)mK3/V. (6.18)
The moment value at which crossover between low moment scaling and extreme
scaling occurs is then given by

rac - ln(V/K 3 )/ 1η(α/κ) (6.19)

Extremely sharp cracks (κ —» 0)


It is seen from equations (6.12) and (6.18) that if κ becomes sufficiently small, a
singular behaviour can occur for low moment values. For the problems described
by equation (6.12) this moment value is

racl = 4, (6.20)

while for the 3-d dielectric problem (see equation (6.18)),

mcl = 3. (6.21)

For m > raci, moments are affected by extremely sharp cracks, and in fact these
moments are singular as κ —> 0.
The calculations of this sub-section show that (for isolated cracks) the
crossover from low moment scaling to extreme moment scaling occurs at rac ~ In L
for K finite. For κ —> 0 however, singular moment scaling may occur as low as raci
as given in equations (6.20) and (6.21). These results allow us to understand in
detail how the conductivity scaling of figure 6.1 crosses over to breakdown scaling
as the importance of the load enhancement at the crack tip increases. With this
intuition we now study the effect of disorder on brittle strength.

6.3 Extreme scaling analysis of dilute networkst 34 ' 38 ' 50 !


In this section, we use an extreme scaling analysis to calculate the strength
of materials containing a small volume fraction, q = 1 — p, of random voids. These
results are compared with the elastic moduli and conductivity of the same models
to illustrate the peculiar scaling behaviour of extreme properties. We emphasize
the theoretical analysis, and defer the experimental implications to section 6.6.
6.3.1 The modelst28'34'38'53'67^
Consider a randomly diluted network whose bonds have a load response
curve such as those given in figure 6.2.
If a small enough external load is applied, the network response is linear.
On application of a larger external load, local failure may begin to occur and the
196 P.M. Duxbury / Breakdown

1.5

0.5 h

0.0
1.5

2.0 I I I I I I I I I I I I I I I I I I
f(d)
1.5

1.0
r
0.5

I . . . . I . . . . I
0.0
0 0.5 1 1.5 2
V

Figure 6.2 Examples of the characteristics of several types of breakdown phenomena: a) the
failure of a spring; b) the failure of a fuse; c) the failure of a dielectric; d) a superconductor
characteristic. In all of these figures, the dotted line is a discontinuous idealization of more
realistic (solid line) characteristics (from ref. 56).
3. Extreme scaling analysis of dilute networks 197

15h

10

0 t 1 1 u 1 1
0 10 20 30 40 50 0.01 0.02 0.03 0.04 0.05
V

a) b)

Figure 6.3 Load response characteristics of random networks: a) a voltage controlled character-
istic for a p = 0.75, 70 x 70 square-lattice fuse network and; b) a strain controlled characteristic
for a p = 0.90, 40 x 40 central-force, triangular spring network (from ref. 53).

network response becomes non-linear. For a sufficiently large external load, damage
grows catastrophically and the whole network fails. Numerical simulation of this
process may be carried out using several different algorithms, such as:
1. Remove the hottest bond at each iteration,
2. Fix the external stress or strain and removing all bonds which have a load greater
than a threshold value.
3. Increase the load at each timestep to simulate a strain rate controlled test.
Most simulations carried out thus far have used algorithm lj 28 l while some
have used 2. Most calculations have been carried out in tensile loading and mode
I failure. Our discussions henceforth will consider only tensile loading conditions,
and the analogous loading conditions in the electrical and dielectric cases.
Two examples of load-response curves calculated for random networks are
shown in figure 6.3 for an electrical problem (random fuse network) and a mechan-
ical (random central-force brittle spring) problem.
In the fuse case (figure 6.3a) catastrophic failure occurs very soon after
the linear regime. This means that very soon after the first bond in the network
fails, the whole network fails by a dramatic crack growth (the network is brittle).
However, if the strain is slowly increased in the spring network, a greater deal
of non-linearity and hence considerable local damage occurs before catastropic
failure. Note however that if the stress were slowly increased as a function of
time, catastrophic failure would occur very soon after the failure of the first bond.
There is thus a great deal of difference between a strain controlled test and stress
controlled test in this problem, and indeed in many fracture problems. The detailed
198 P.M. Duxbury / Breakdown

Figure 6.4 a) A random spring configuration (p = 0.90) before application of the external
load; b) the configuration at loss of rigidity and; c) the configuration after complete fracture has
occurred (from ref. 53). ε is the measured strain.

microstructural process of failure in the spring network is illustrated in figure 6.4.


The middle configuration corresponds to the strain at which the spring
network looses rigidity without disconnecting due to the rotational freedom of the
central force springs. At this point there is a band of bonds with zero shear modulus
or a special sort of shear band. Note that this shear band is able to flow only a
maximum of one lattice spacing before the shear modulus is re-established. Final
failure occurs very close to the shear band.
As the initial disorder increases, the fracture path becomes more irregular
as it tries to take advantage of pre-existing voids. This is illustrated in figure 6.5
for the electrical case.
As q —> qc, the fracture surface (path in 2 dimensions) is expected to
become fractal on length scales L < £, although this fractal dimension has not
yet been quantified. It would be useful if more numerical study of the structure
of the fracture surfaces of idealized models were carried out for comparison with
the results of fractographic studies on real materials (for further discussion of this
3. Extreme scaling analysis of dilute networks 199

Figure 6.5 Fracture paths in square lattice random fuse networks: a) p = 0.90 and; b) p = 0.70
(from ref. 38).

issue see section 6.6.3).


In the remainder of this chapter, we emphasize studies of the strength of
the random brittle networks as a function of disorder and system size. The issue
of complex crack topologies is discussed in more detail in chapters 8 and 9.
6.3.2 The average strength of dilute brittle networks
We calculate the failure stress and breakdown current of dilute networks, by
noting (see figure 6.3) that in these loading conditions, these networks are brittle.
The breakdown stress (or current) is then proportional to the reciprocal of the stress
(or current) intensity at a "hotspot" in the network. Due to the lack of non-linear
deformation before failure, we can calculate the local field at these hotspots using
linear theories. The essentially new part of this analysis is to combine the above
linear analysis with a statistical prediction for the typical size of these hotpots
as function of disorder. These problems have not yielded to a rigorous analysis,
though as we describe below they do allow a convincing analytic analysis that is
intuitively appealing and which agrees with numerical simulations.
The analysis proceeds by constructing approximate upper and lower bounds
to field hotspots by considering various defect configurations that lead to strong
load enhancements. The first such configuration we study is a crack-like sequence
of voids in an otherwize pure environment. The typical size, a, of the largest such
crack-like flaw aggregate is estimated by the "extreme scaling" argument

Vqa ~ 1, which implies a ~ - In V/\n q (6.22)


200 P.M. Duxbury / Breakdown

The factor Vqa is a product of the probability of finding a adjacent removed bonds
on the lattice, times the number of places in the network at which such a con-
figuration could be located. It is assumed that the surroundings of this crack-like
object do not alter the probability significantly. This is an accurate approximation
in the pure limit (small q) where the average number of voids in the immediate
neighbourhood of the crack is small, but must be altered as q —> qc. The stress
intensity at the tip of this crack-like (lattice) flaw scales as

σύρ/σ0 ~ 1 + ka1/2. (6.23)

Here k is an unknown constant. This is a little different than the stress intensity
near the end of a straight continuum crack (see equation (6.10)) by virtue of the
fact that the lattice cuts off the 1/r1/2 singularity. In real materials, this cut-off is
naturally provided by the plastic radius, or by other local mechanisms (damage,
or decohesion) which take place at a well defined small distance near the crack tip.
The additive factor 1 in equation (6.23) is included to ensure that as a —> 0 the
correct pure limit bond strength is recovered. Note that although the form (6.23)
is correct in both the a —> 0 and a —> oo limits, it is not exact on the whole a
regimeJ60^ It does however provide a useful interpolation formula between the two
limiting cases. Some alternative forms (for the electrical case) have been recently
suggested J73l
If failure initiates at an isolated crack-like flaw, catastrophic failure will cer-
tainly occur. However local failure may occur at lower external loads, for example
at a thin neck between two crack-like flaws, w h e r e ^

^tip/^o ~ 1 + ka. (6-24)

Here k is an unknown constant. For a crack-like flaw with finite sharpness (which
applies to all lattice cracks), the stress enhancement at the crack tip can scale at
most as a, just due to conservation of field lines. Long range cumulative effects due
to many distant flaws could conceivably lead to stronger hotspots. A dipole argu-
ment shows that this is not the case for dilute networksJ50^ Combining equations
(6.22)-(6.24) we then deduce

ac{q)/ac(0) ~ l/[l + km{-\nV/\nq)a} for q < qc (6.25)

Here km is an undetermined constant and l/[2(d — 1)] < a < 1. The lower bound
on a is from the d-dimensional crack-like flaws of the type leading to equation
(6.23), while the upper bound is from equation (6.24).
The simple arguments leading to equation (6.25) are very general and ap-
ply to conceptually similar problems in electrical and dielectric breakdown. In the
electrical case the burnout current of a random network with bonds having a char-
acteristic such as shown in figure 6.2a (fuse network) is found to behave asi^34,3^

Jc(g)/Jc(0)~l/[l + M-lnV71n<zn for q < qc. (6.26)


3. Extreme scaling analysis of dilute networks 201

1.0

0.6

0.2

Figure 6.6 a) The conductivity (+) and breakdown current (v) of resistor networks, calculated
from 50 realisations of 100 x 100 square lattices (from ref. 56) and; b) the elastic modulus (x)
(from ref. 78) and tensile strength (Δ) (from ref. 53) of spring networks (from 50 x 50 lattices).

The analogous problem of dielectric breakdown in insulators containing metal in-


clusions behaves as^34'52!

£c(q)/Sc(0)~l/[l + kd(-\nV/lnqf] for q < qc (6.27)

where 1/2 < ß < 1. The different dimensional dependence in the lower bound in
the dielectric case is due to the fact that one dimensional conducting paths can
cause catastrophic failure in all dimensions for that problem, while in the fuse and
mechanical cases, an Ld~x-dimensional fracture surface is necessary.
The formulae (6.25)-(6.27) show two striking behaviours:
1. A size effect. For fixed q < qc the network strength reduces logarithmically with
system size.
2. A dilute limit singularity. The network strength is non-analytic (in the thermo-
dynamic limit) on approach to q = 0.
The behaviours 1 and 2 are markedly different than the scaling behaviours
of transport or elastic moduli in the same limit. We show this difference explicitly
in figure 6.6 where numerical simulations for the conductivity, elastic moduli and
strengths of two-dimensional networks are displayed.
It is seen from these figures that even on 100 x 100 and 50 x 50 arrays the
dilute limit singularity in strength is strong. Due to the logarithmic size effect this
effect becomes stronger with increasing system size, and approaches a jump singu-
larity in the thermodynamic limit. The numerical curves are in excellent agreement
with the theory of equations (6.25) and (6.26). Experimental systems showing a
behaviour similar to that of figure 6.6 will be discussed in sections 6.6.1 and 6.6.2.
202 P.M. Duxbury / Breakdown

6.3.3 Strength fluctuations^38'40'56!


It is well known that average moduli vary over lengths L < £, but saturate
on lengths L > £, where £ is the percolation correlation length (see e.g. Stauffer's
book in ref. 1). Since regions on lengths L > ξ are essentially uncorrelated, the
average value of moduli measured on samples of size L have fluctuations of order
(L/£)~ d / 2 . This comes from arguing that the modulus measured on a system of
volume, V, is an average over {L/£)d independent values of the modulus in volumes
of approximate size, £d. The distribution of sample to sample variations in moduli
then obeys the familiar central-limit form
P(Xv)^exi>[-V{Xv-X00)2/c(p)} for L > £, (6.28)
where P(Xv) is the probability that the modulus measured on a sample of size
V has a value Xv, and Xoo is its value in the thermodynamic limit. c(p) ~ c£3 is
independent of V for large V. Since the distribution in (6.28) is very narrow, sample
to sample fluctuations are also small for L > £. This is the origin of the statement
that, conductivity and elastic moduli self-average for large enough system sizes.
Because breakdown and other extreme properties^ are dominated by un-
likely events, the discussion above no longer applies. In these cases the distribution
of fluctuations is broad and asymmetric (non-Gaussian). This is demonstrated
numerically for the electrical case in figure 6.7, where we have calculated the dis-
tribution function for the conductivity and the largest current in a dilute electrical
network.
It is possible to derive the form to be expected for the distribution given
in figure 6.7b in several ways depending on the amount of mathematical rigor you
wish to pursue. The most direct one is to use the fact that the distribution of
bond stresses is exponential for large stress (for q <C qc)-^ Then the cumulative
probability, C\(a < a m a x ), that a bond stress is less than a maximum value, a max ,
is given by
<7max

Cx{a < amax) ~ j exp{-ba')da' (6.29)


o
where b is assumed independent of system size. It was seen in section 6.3.2, that
large local stresses are induced by special local defect configurations. Since the
defect configurations are random and uncorrelated, it is reasonable to assume that
the large stresses lying in the tail of the bond stress distribution are uncorrelated
(this appears to work for dilute networks, but should be kept in mind when trying
to extend these ideas to other problems). Assuming the bond stresses to be un-
correlated, the probability that no stress in the V = Ld bonds in the network is
greater than a max is given by
Λ^(σ < a m a x ) - [1 - exp(-bamax)/b]v (6.30)
Since am3LX is large compared to 1/6, this expression is approximately
RvW < ^max) ~ exp [-{V/b) exp(-ba max )] (6.31)
3. Extreme scaling analysis of dilute networks 203

0.3 0.15

*
0.2 h
Λ
O i ° i
\
V
1" ° i
Ü
* 0. 1 h

I ■ ' ■ » I « ■ ■ j>1 A . . . I . . . . i
0.0
0.5 0.75 1.26 1.5
G/<G> W/<Ia»x>
(a) (b)

Figure 6.7 The probability that a 50 x 50 random resistor network (RRN) will: a) have con-
ductance G and; b) contain a maximum current J max in any bond. These distributions were
calculated from 2000 realizations of the RRN at q = 0.10. (G) and (7 max ) are the average values
of the conductivity and maximum current respectively (from ref. 56).

The differential probability that a system of size V has maximum stress a max is
given by;
iM^max) ~ ^ e x p ( - 6 a m a x ) e x p [ - ( F / 6 ) e x p ( - ^ m a x ) ] (6.32)
This form is clearly skew, and the analogous form for the maximum current (replace
0"max by /max) provides a good fit to the data of figure 6.7b.
Using similar arguments to that described above for the distribution of
the largest stress, we may calculate the reliability of these dilute networks. The
reliability of a network at a given external load is the probability that the network
will fail at or below that load. To calculate this cumulative probability for dilute
networks, we first note that since the distribution of large (vacant) clusters in
the percolation problem (with q < qc) is exponential,^ the cumulative probability
that a crack has size a is given by a form identical to equation (6.31) with the stress
replaced by a, the crack size. Combining this form with the expected behaviour of
the averages (6.25)-(6.27), we then deduce that the required reliability distribution
for the tensile strength of porous brittle materials is[34>38>40>56]

Cv(ac) ~ 1 - exp [-kVβχρ(-ξι9^α)] , (6.33)

where
6 = - Hi) and gm = [<TC(0) - σ^)]/[^σ0(ς)} (6.34)
204 P.M. Duxbury / Breakdown

andfc,and km are unknown constants. A similar analysis may be carried through


for the electrical and dielectric case, so that

Cv(jc) ~ 1 - exp [-W exp(-6<7 e 1/a )] (6.35)

where
9e = \jc(0)-Mq)]/[ke3c{q)]. (6-36)
For the dielectric case, we find

Cv(£c) ~ 1 - exp {-Wexpi-t^0)} , (6.37)

where
9d = &(0) - Sc(q)]/[kd£e(q)] (6.38)
Now note that if a = 1 or /? = 1 these reliability forms further reduce to

CV(XC) ~ 1 - exp [-cVexp(-khX0/Xc)] (6.39)

where c,feare functions of g and equation (6.39) applies to electrical, mechanical or


dielectric failure with X = σ, 5 or j . This form may be found also be found directly
from (6.31) by using the substitution ac(a)/ac(0) « cro/^max and Cy = 1 — Ry-
The distribution (6.39) is different from the usual Gumbel form (C(X) = 1 —
exp[—cexp(fcX)]) used in fracture reliability analysisJ75^ The usual Gumbel form
is based on a weakest link argument that implicitly assumes that all parts of the
material carry the same stress. The analysis leading to equation (6.39) assumes that
all surviving parts of the material have the same strength (so there are no weak
links in the usual sense), and that failure is initiated by load hotspots. Equation
(6.39) is thus of an extreme maximum form rather than the extreme minimum form
of the usual Qumbel distribution.
The reliability distribution of dilute networks may be generated numer-
ically by simulating a large number of realizations of random networks. These
calculations, carried out for two-dimensional fuse networks are displayed in figure
6.8.
It is seen from figure 6.8 that the reliability distribution (6.39) provides a
good fit to the fuse network data. Most experimental reliability data is however
fitted to the Weibull form:

C K ) = 1 - exp[V(ac/as)m}, (6.40)

where m is the Weibull modulus and as is the Weibull scale factor. Equation (6.40)
also appears to give a good fit to the data of figure 6.8a, so a more sensitive test
between the two possible forms is necessary. We thus define the form

A = \n(-\n[l-C(ac)]/V). (6.41)
3. Extreme scaling analysis of dilute networks 205

1.0

C(tf c )

b)

Figure 6.8 The probability, C ( X ) , that the first bond in a fuse network fails when: a) an external
voltage of size V\ is applied to it (from 1500 configurations of 50 x 50 square networks at q = 0.10)
(from ref. 38) and; b) a stress ac is applied to it (from 1000 configurations of 30 x 30 triangular
networks at q — 0.10) (from ref. 54). The solid line in a) is a fit to either of equations (6.39) or
(6.40).
206 P.M. Duxbury / Breakdown

0.0Lf$- __l L_
0.5 0.7 0.9 1.2
IniUV,)

15.0

10.0

o.o 4j—-^
1.5 2.0 2.5 3.0
b)

Figure 6.9 Unbiased tests of reliability distributions using A as defined in equation (6.41) and
the data of figure 6.8a: a) the Weibull form (6.40); b) the modified Gumbel form (6.39) (from
ref. 38).
4. Fractal and hierarchical microgeometries 207

If equation (6.39) is correct a plot of A against σο(0)/σ€(ς) should be linear. If the


Weibull form is correct, a plot of A against 1η[σ€(#)] should be linear. Unbiased
tests of these forms are presented in figure 6.9.
It is seen from figure 6.9 that the form (6.39) provides an improvement over
the Weibull distribution for dilute fuse network reliability. A similar conclusion has
been drawn for the dilute random central-force spring caseJ53,54l Extending these
conclusions to three dimensions, we propose equation (6.39) as an alternative form
for the reliability of brittle systems with random dilution. The broader implications
of this result for the reliability analysis of materials is discussed in section 6.6.4.
As seen in figure 6.7, extreme properties have stronger sample to sample
fluctuations than low moment properties. From equation (6.39) we may estimate
the expected sample to sample variations in strength by comparing the location of
the 25th or 75th percentiles in figure 6.9 with the 50th percentile - the "average"
value. Defining the strength fluctuations a s ^

Δσ 0 « [σ(750ι) - σ ( 5 0 Λ ) ] / σ ( 5 0 Λ ) (6.42)

we find that for In V >· £i


Aac ~[K{p)/hiV]. (6.43)
Strength fluctuations in dilute networks thus reduce very slowly with system size,
in contrast to conductivity or elastic moduli fluctuations which have sample to sam-
ple variations 0(1/\/V). A similar analysis may be carried out for the dielectric
and electrical cases and it again shows that their strength fluctuations are loga-
rithmic for I n F ^> £χ. For p approaching the pure limit, the strength fluctuations
become smaller, so that exactly at the pure limit, they are of course zero. A prac-
tical consequence of the large sample to sample fluctuations occurring in measured
strength values, in combination with the destructive character of typical experi-
ments, make experimental studies of breakdown phenomena more demanding than
corresponding experiments on transport or elastic moduli.

6.4 Fractal and hierarchical microgeometries


6.4.1 Strength scaling near the percolation point^34,45-47,56^
Within the nodes, links and blobs picture (see Stauffer's book in ref. 1), as
ξ becomes large we may visualize the percolation backbone as a lattice of nodes
with lattice spacing £, between which are links decorated by blobs. The node lattice
is considered to be regular, and the greater effective lattice constant (~ ξ) of the
renormalized lattice leads to an electrical strength which reduces as^34,38,56]

Jc(<z) ~ j c ( O ) / ^ - 1 ~ jc(0){qe - q)(d-l)v as q - qc. (6.44)

Similarly for dielectric problems, we find'34'38'56'

5C(«) ~ £c(0)/£ ~ £c(0)(«c - «)" as q^qc. (6.45)


208 P.M. Duxbury / Breakdown

In mechanical problems, the situation is more complex, as, due to a cantilever


effect, the moment acting on a given link of the network must be considered. An
upper bound on the fracture stress is found by assuming that this "moment arm"
is of order unity, so that^46'56!

σΜ) < *c(0)(fc - q){d-1)v- (6-46)


A lower bound on the network strength is found by assuming the moment arm is
of size £, so that

*c(«) > *c(0)(«c - q)(d-1)vH ~ *c(0).(ft - qf (6.47)


46,5
Critical strength scaling in mechanical problems then behaves as^ ^

<Tc(q) ~ ^ c (0)(g c - q)x as q -► qc (6.48)


with
(d- l)i/ <x < du. (6.49)
These arguments apply equally well to central-force networks and elastic models
with bending forces. From equation (6.44), it is straightforward to derive the break-
down electric field critical behaviour of the fuse network by including the scaling
behaviour of the network conductivity (see chapter 4)

Sc(q) ~ €e(0)(qe - qf-1»-* as q - qc, (6.50)

where t is the conductivity exponent. Similarly, the breakdown strain behaviour of


the elastic networks is given by equation (6.48) with a correction due to the critical
scaling behaviour of the elastic modulus (see chapter 4)

£c(q) ~ ec{0)(qc - q)x~T as q^> qc, (6.51)

where T is the elasticity exponent. It is interesting to note that the breakdown


electric field and strain may either diverge or go to zero near qc depending on the
relative size of the two exponents v and t in equation (6.50) and x and T in equation
(6.51). If £ c or ec diverge on approach to qc, their behaviour as a function of q is non-
monotonic. That is, they decrease for small q due to dilute limit scaling (equations
(6.25) and (6.26)) and diverge for q near qc due to critical scaling (equations (6.50)
and (6.51)). This non-monotonic behaviour has been observed in fuse and bubble
(see section 6.4.2) models.[38'42]
It is possible to extend the critical scaling arguments discussed above to
continuum percolation problems, where each link itself is composed of sublinks
(necks) of various strength. The strength of a model then depends on the distri-
bution of sublink strengths in addition to the long distance percolation geometry
characterized by £. This sort of argument has been used to calculate the break-
down critical exponents relevant to electrical, dielectric and mechanical problems
in "swiss cheese" and "inverted swiss cheese" geometriesJ47»57!
4. Fractal and hierarchical microgeometries 209

It is possible to combine the dilute limit scaling laws given in equations


(6.25) to (6.27), with the critical scaling laws of equations (6.44), (6.45) and (6.48).
This is achieved by replacing L in equations (6.26) and (6.27) by L/ξ. This sub-
stitution in equation (6.26) y i e l d s ^

jc(q)/jc(0) ~ Γ ( < , - υ / [ 1 + * , ( l n ( £ / i ) / £ i H . (6-52)

This formula interpolates smoothly between the two singular behaviours in the
limits q —> qc and q —» 0, and provides a good approximation to strength be-
haviour over the whole dilution range. A similar extension of equation (6.27) for
the dielectric breakdown problem leads to

€c(q)/€e(0) ~ r X / [ l + Μ 1 η ( £ / 0 / ω " ] · (6-53)

The mechanical fracture problem requires a slight modification of this procedure


to yield,
'c(*)/*c(0) ~ Γ β / 7 [ 1 + M l n ( L / i ) / 6 ) e ] (6-54)

6.4.2 Chain of bundles (bubble) modelst6"23'39'42'72!


Chain of bundle models initiated with studies of the strength of textile
yarns^l (for example cotton or wool). The fibers constituting the yarn are held
together by friction. Over lengths / < 6, the "ineffective length", the fibers may
slide. While for I > 5, the fibers are held together by friction forces. For the purposes
of modeling, the yarn is then divided into a sequence of fiber bundles of length δ.
These fiber bundle models have also been used to model (continuous) fiber
reinforced materials, in which case a thin fiber laminate is of the form shown in
figure 6.10. An effective stress transfer length, again denoted δ is defined as the
distance over which the matrix transfers stress to the fiber.
One essential assumption of the fiber bundle models is that each fiber bun-
dle may be treated independently of the other fiber bundles, and that load sharing
only occurs amongst fibers in the same bundle (this is in contrast to the arrays
considered in sections 6.3 and 6.4.1 which allow load sharing in all directions). On
application of a tensile stress to the yarn or laminate (matrix and interface effects
are only included via effective fiber properties, and local load sharing rules), weak
fibers fail and the load that they previously carried is shared by intact fibers in
the same fiber bundle. The load sharing follows "ad-hoc" rules such as: equal load
sharing, where all neighbouring intact fibers share the load equally and; local load
sharing, where a specified number of near neighbours share the load. The equal
load sharing case is appropriate to loosely wound yarn, while the latter applies to
the composite case where local fiber failure leads to strong stress enhancements in
neighbouring fibers.
Given a cumulative probability of failure for a particular fiber element,
Ci(ac) and a load sharing rule, we wish to find the cumulative probability of
210 P.M. Duxbury / Breakdown

fiber -*
element
(bond)

£| I I I I I I I I I I I
< w >

Figure 6.10 Schematic drawing of a thin fiber laminate, along with definitions of the parameters
of the fiber bundle models. Under applied loading, each fiber bundle is assumed to carry the same
load.

failure of the chain of fiber bundles (RLW(&C))I where L and W are as defined
in figure 6.10). There is a reduction in complexity due to the one-dimensional
structure of the model, as the strength of the weakest fiber bundle determines the
strength of the chain. Thus if we define Cw{^c) to be the probability that a fiber
bundle of width W fails at a stress below ac, then

RLW^C) = 1 - [1 - Cw(ac)]L . (6.55)

The calculation of Cw(^c) is however far from trivial. Direct application of ex-
treme distribution theory (for identically distributed independent random vari-
ables) to a fiber bundle is not obviously correct as the fiber elements (bonds) fail
by a correlated damage evolution process. Much of the work in the mathematics
and engineering communities^ 6-23 ' 3 ^ has concentrated on fiber bundle models with
bonds having continuous failure distributions and either local or equal load sharing
damage evolution algorithms.
Fiber bundle models with random dilution^42,72! (also called bubble models)
were introduced as a soluble approximation to the percolation problem on regular
lattices. There, equal and local load sharing rules have been used and in the most
general case, the bonds are allowed to have two different moduli. A summary of
some of the main physical results evolving from fiber bundle models are as follows;
1. Models with Weibull distributed bond failure and equal load sharing lead to a
normal distribution in C\y and a stable limiting strength (there is no size effect in
average strength for large system sizes)S^
2. Models with Weibull distributed bond failure and local load sharing lead to a
Weibull failure probability (with a variable modulus) and exhibit a size effect in
average strengthJ 17 '
4. Fractal and hierarchical microgeometries 211

3. Models with diluted fiber bundles, equal load sharing and an exponential length
to width ratio exhibit a size effect in strength very similar to that found in the
dilute networks discussed in section 6.3J42^
The bubble models are attractive theoretically in that they are often an-
alytically tractable and show many of the interesting strength scaling features of
more complex models. It should be emphasized however, that these models (at
present) only allow planar crack topologies, a limitation that must be carefully
considered in using these models on random systems. Nevertheless they are an
important theoretical tool in the study of breakdown phenomena.

6.4.3 Cayley trees and hierarchical bubble models


A Cayley tree hierarchical microgeometry is shown in figure 6.11. Turcotte
et α/f24'25^ studied a model where the bonds of a Cayley tree have Weibull dis-
tributed failure probability, and using a load sharing rule where a failed bond
shares its load with all intact bonds on the same level and same branch of the tree.
If no such intact bond exists, the load is shared with the bonds on the branch which
is closest to the failed branch. An approximate renormalization-group analysis (see
also chapter 5 of this book) using as a hypothesis that the Weibull distribution is a
stable limiting distribution for the strength of the tree, led to the conclusion that
the tree had a stable limiting strength. It has been recently shown rigorously,^
that the latter conclusion is incorrect and that the strength of the tree decreases
as l / l n ( n ) . Using the usual conversion to an effective euclidean space problem,
(L = 2 n ), the size effect in Cayley trees is 0 ( l / l n ( l n L ) ) . This is a very slow size
effect in strength and for most purposes can be considered to be equivalent to a sta-
ble limiting strength. Note that this size effect is much slower than the 0 ( 1 / In V)
size effect of dilute regular networks (see section 6.3).
The Cayley tree microgeometry has also been used to study dielectric
strength in random mediaJ33^ There the dielectric strength of insulating Cayley
trees containing randomly placed metal bonds was calculated by noting that the
path along which the minimum number of removed bonds lie provides an estimate
of the tree's strength. This "minimum insulating gap", #, is finite for q < qc and
has a critical scaling g ~ (qc — q)u'. The minimum gap scaling near the percolation

Figure 6.11 A n a = 2 (a = the number of branches) Cayley tree of n = 3 generations.


212 P.M. Duxbury / Breakdown

Figure 6.12 Model for hierarchical porous material

point is identical to the critical scaling behaviour of the dielectric breakdown field
strength given in equation (6.45).
6.4.4 Hierarchical porous materials
There have been some recent attempts to modify the Griffith formula for
the stability of a cleavage crack in an elastic medium, to the case of inhomogeneous
mediaJ31»36! The most general of these is by C o o k ^ who has presented an analysis
of a porous material with the microgeometry shown in figure 6.12. Building on
earlier papersJ31»36! the scaling behaviour of the strength and toughness of this
hierarchical system was calculated using energy balance arguments in combination
with plausible forms for the scaling behaviour of the surface and elastic energies. As
usual, the strain energy release rate, Um was balanced against the energy to create
new fracture surfaces U8. In the simplest Griffith theory,^ the crack driving force
G\ = —dllm/da « Ε0σ02α for uniaxial stress loading conditions, and the crack
resistance R = dU8/da « 7, where 7 is the crack surface energy. In the hierarchical
geometry of figure 6.12, it was assumed that

R ~ aD~3 (6.56)

and
Gi - σ ο ν , (6.57)
where y is an undetermined exponent of the theory and D is the mass fractal
dimension of the hierarchical structure (e.g. of figure 6.12). Equating R and G\
leads to the fracture stress:
ac ~ a ( D - 3 - ^ 2 . (6.58)
For the standard Griffith case D — 3 and y = 1, and the standard square root
reduction in strength with crack size is reproduced. In the case of hierarchical
porous material, D < 3 and we expect (although this is on less solid ground)
y ~ 1, so that strength reduces more rapidly than in the Griffith case.
5. Toughness of disordered materials 213

On short distances (small crack sizes) there is often an additional contri-


bution to the strain energy release rate which is of the form

G2 ~ a~9, (6.59)

which decreases with increasing crack size. This term may arise for example from
residual stresses frozen into the material during fabrication,^ or in the case of
indentation or notch testing due to residual stresses produced during production
of the controlled crack initiation site. The inclusion of the crack driving force of
equation (6.59) to that of equation (6.57) (which increases with crack length), leads
to the possibility of crack arrest. In this case an initially unstable crack may grow
under the influence of the driving force (6.59), and then arrest at an intermediate
crack size where the sum of the forces from (6.59) and (6.57) is a minimum. In the
ceramics community, materials which exhibit this effect are said to show "R or T"
curve behaviour J79^
6.4.5 Material containing a fractal crack
As will be described in chapter 8, there has been considerable study of
mechanisms for the generation of fractal cracks in both dielectric^80-83! and me-
chanical systemsJ 84-88 ! Tree-like structures are also observed experimentally in
both sub-critical and high speed breakdown processes. In the former case, it is
of interest to determine the effect of fractal cracks on material properties, and
whether it is possible to detect fractal cracks by non-destructive means. There has
been little study of this issue,[87'891 and here we illustrate the problem by discussing
a calculation of the effect of a fractal conducting crack on the dielectric strength
and capacitance of an insulator J89^
We assume a two-dimensional insulator in a plate geometry, and grow a
fractal crack into it using the DLA algorithm. This crack structure may for example
model the growth of a corrosion crack into the insulation of a submarine cable. The
sea water makes the tree-like crack conducting and changes the capacitance and
dielectric strength of the cable. Numerical simulations of this effect for the fractal
crack of figure 6.13a are shown in figure 6.13b.
It is seen from figure 6.13b that a small fractal crack dramatically reduces
the dielectric breakdown voltage of the insulator, and that capacitance measure-
ments provide little indication of the presence of such a crack. This behaviour is
similar to that observed for cleavage cracks (see figure 6.1), although the precise
functional form of the strength and capacitance perturbations are expected to be
different than those for straight cracks.

6.5 Toughness of disordered materials


Mechanical stength is obviously a very important property in the engineer-
ing applications of materials. However, a property of equal importance in many
applications is the toughness of the materialJ4,76,79^ Toughness is a quantity that
214 P.M. Duxbury / Breakdown

KPL·
"U

iA I I

a)

1.0
R>J 1 , ( , | ■ i — | r

0.8 h \
Γ" V ^ Λ\ m=4 J
1/C
V \s ■i\/S«

0.6
Vji-10 ^N^X ]
sm=15 ^ ν ^ N/V
0.4 h
I ν
ΛΑ.
^Α^ J
0.2 h

1
i. 1 i I · 1 ^ Π
0.2 0.4 0.6 0.8 1.0

b) N b /N t

Figure 6.13 a)Structure of a fractal conducting crack in an insulating background and; b) the
effect this crack on the capacitance, C, the dielectric breakdown voltage, Vh and voltage moments,
with m = 4,10,15, as a function of crack size (measured by the number of bonds on the crack
iVb. Nt is the number of bonds on the final crack pattern that connects the two plates together.)
5. Toughness of disordered materials 215

measures the amount of energy that a material can absorb before fracture. It is
a good measure of impact resistance, flaw tolerance, and durability. Many metals
are mechanically tough as a large amount of energy can be absorbed by dislo-
cation emission before catastrophic fracture occurs. This type of toughening can
be well approximated by studying an expanded zone near the crack tip, in which
dislocation pile up occurs. With the development of high strength fiber re-inforced
composites that do not have significant toughening by dislocation emission, it has
become necessary to develop novel toughening mechanisms, and many such mech-
anisms have been suggested. A good fraction of these mechanisms rely on stopping
or deflecting a growing crack by the use of strong obstacles, and this sort of pro-
cess is especially relevant in composites such as fiber reinforced composites and
polymers containing rubber inclusions. In such a mechanism, a crack may start
growing, and then be stopped by an obstacle (e.g. a fiber). If the external stress is
further increased, a crack may initiate in an entirely different part of the sample.
This is a non-local toughening process which cannot be studied by single crack ar-
guments alone, and has so far evaded a convincing analytic analysis. Some progress
has recently been made using network models and numerical modelling, and we
now discuss those models and results.
As stated in section 6.3, the simplest diluted networks are often brittle,
as once a crack has begun to propagate at constant applied stress (mechanical
breakdown), current (electrical breakdown) or electric field (dielectric breakdown),
it usually leads to the failure of the whole network. Toughening of random networks
has been studied using electrical and mechanical networks in two dimensions. Two
mechanisms have been considered:
- Toughening due to distributions of bond strengths J32'60! This is essentially a crack
deflection mechanism, as strong bonds lie in the path of any advancing crack, and
cause it to change its propagation path locally. Numerical calculations and analytic
arguments suggest that although this mechanism causes significant toughening (or
ductility in the language of refs. 32 and 60) for smaller system sizes, a transition to
brittle fracture appears to occur if the system is made large enough. These models
will be discussed in detail in chapter 7.
- Toughening due to bonding behind the crack tip (residual bonding)J51^ As well
as being of importance in the toughening of ceramics^4,91'92! this mechanism may
occur in fiber-reinforced composites undergoing matrix fracture with fiber pullout.
This is illustrated in figure 6.14, where the matrix crack is bridged by fibers that
have undergone pullout, but which still provide a force holding the crack together.

We have generalised the dilute network models to include this effect, by


assigning the bonds behind the crack tip a small (residual) elastic constant. The
amount of residual bonding is quantified by the ratio
R — residual bonding parameter = E0/ET, (6.60)
where E0 is the background elastic constant, and Ex is the elastic constant of the
216 P.M. Duxbury / Breakdown

Figure 6.14 Schematic example of a composite failing by fiber pullout. The unbroken fibers
behind the crack tip reduce the stress intensity factor at the crack tip (from ref. 51).

residual bonds. The effect of these residual bonds on the stress intensity at the tip
of the crack is shown in figure 6.15.
In this calculation, a crack (line of removed bonds) of initial size 6 bonds
was placed in a 40 x 40 central force network, and an external tensile stress applied
perpendicular to the long axis of the crack. Bonds change irreversibly from E0 to
ET when a critical stress occurs across them. It is seen from figure 6.15, that after
an initial increase in stress intensity, the residual bonding behind the crack tip
carries enough of the applied load to actually reduce the stress intensity at the
crack tip. A fixed applied stress that is sufficient to initiate crack growth, is then
not sufficient to cause catastrophic failure, and this system shows crack arrest.
A similar effect may be produced in electrical networks, where the analog of the
residual elasticity is a (small) residual conductivity that remains after a bond has
failed electrically. This electrical problem is more tractable numerically than the
mechanical networks, and provides a good pedagogical model in which to study
this toughening mechanism. We now discuss the toughening transition in electrical
networks with residual conductivity.
A quantitative measure of toughening in these problems, is the amount of
damage that occurs in the network before catastrophic failure. In these random
networks, the damage is measured by the number of bonds that break before final
5. Toughness of disordered materials 217

1.Θ

1.6 h

1.4 h

1.2

Figure 6.15 The stress intensity at the tip, σ<;ρ, of a crack of length, a, which has residual
bonding, R = 20, behind the crack tip: (□) for an initial crack of size 6, σ ί ι ρ falls below its initial
value at length a^; (Δ) for an initial crack of size 8, σ ίιρ falls below its initial value at length ag
(from ref. 51).

failure. For brittle problems, we expect most of the failure to occur at the fracture
surface, and for material with random voids (and q <C <jc), this implies:

or nb = Nb/Ld x
& constant for brittle fracture. (6.61)

For highly toughened networks, considerable damage occurs throughout the mate-
rial so that
nb ~ Ly with y > 0 (6.62)
y is the toughening exponent, and if y > 0, the random material is toughened. nb
is an "order parameter" for the transition from brittle to tough networks. A plot
of nb against R is shown in figure 6.16 for p = 0.75.
As R —> oo, equation (6.62) holds and the network is brittle. For JR small
however, n^ increases strongly with system size, indicating a toughening trend, and
y > 0 as indicated in equation (6.62). An interesting feature of figure 6.16 is that
there appears to be a fairly well defined transition between the brittle and tough
regimes at R « 35, so that for R > Rc the network is brittle and for R < Rc
it is tough. We anticipate that a similar analysis extends to mechanical systems,
and that the existence of such critical toughening parameters is of considerable
practical as well as its fundamental interest.
218 P.M. Duxbury / Breakdown

1 1 I I I I I I , I I ,

* o

2h
LD »
_J

o
55 D
D

lb
k O
o
D
o
O
9 β 9 » 1

III

10 100 1000
R

Figure 6.16 The number of bonds broken in the network failure, iVb/L, as a function of the
residual resistance ratio R (from ref. 51). p — 0.75 and each point represents an average over 20
configurations of square lattices of size L — 10 (<>), L = 20 (□), L = 30 (*), and L = 40 (o).
6. Discussion and experimental implications 219

The brittle to tough transition may be studied using fiber bundle models
such as those discussed in section 6.4.2. In general equal load sharing models are
expected to be tough, while local load sharing models tend to be more brittle.
There are of course exceptions to this rule, and it has been recently shown that
the diluted bubble model, with two values of bond conductance, shows a brittle to
tough transition at R = 4J72^ Note however, that the criterion for this transition
is different than that used in figure 6.16.

6.6 Discussion and experimental implications


6.6.1 The tensile strength of porous materials
Kendall^92! reviews work on the tensile strength of porous materials and
suggests as a useful relationship

ac(q, a) = [(E0RO{1 - 1.9q + 0.9q2) exp(-dq)) /(πα)] ^ , (6.63)

where R0 is the fracture energy of the dense material, d is the grain diameter and
a is the size of the dominant crack (essentially a free parameter in the theory). He
thus considers that strength is a function of two independent parameters q and a.
Now note that equation (6.25) may be adapted to predict of the effect of random
pores on strength if we make the association

V — Ld ~ (volume of sample)/(average volume of a pore). (6.64)

For systems with a narrow distribution of closed almost spherical pores. Four im-
portant predictions of equation (6.25), with the substitution (6.64), which differ
from (6.63) are as follows:
1. Strength depends only on q - the dominant crack-like pore a aggregate being a
function of q and not independent of it as in (6.63).
2. A size effect - the size effect is logarithmic for small L > ξ and algebraic (critical
scaling) for L < £. The size effect in strength has been studied intensively both
experimentally and theoretically, and Harter provides a list of 700 references to the
problem in a 1977 reviewJ104^ The novel feature of the work described in this book
is the correlation between a particular microstructure and a specific size effect.
Experiments which address this correlation for a range of well controlled random
microstructures are sorely needed.
3. A dilute limit singularity - there is a rapid drop in strength at small pore vol-
umes. This effect is not well documented, although we reproduce in figure 6.17 data
showing this effect. The rapid drop in strength occurs at very small pore volume
fractions, and for this reason has probably been excluded from most studies of the
strength of porous materials.
4. Critical scaling near qc - when a material is very porous it begins to approach
zero intrinsic strength. There have been a couple of experiments on idealized two-
dimensional systems which are consistent with the theory,t36'45! although (as dis-
220 P.M. Duxbury / Breakdown

I 1 1 1 1 1

90
l I
*<—-Pure System

σ 5
° Tii 1
, (MPa)

30L·
\ u fi

M T A

0I 1 1 1 1 1
0 0.1 0.2 0.3 0.4 0.5
q
Figure 6.17 The strength (uniaxial tension) of a porous boro-silicate glass as a function of pore
volume fraction. The pores are spherical with diameter 60 μπι (from p.367 of ref. 93).

cussed in section 6.6.1), there is room for more sophisticated experimental studies
of critical strength scaling in porous materials.
From the discussion in 1-4 above it is seen that the new theory of the
strength of porous brittle materials embodied in equation (6.25) with the identi-
fication (6.64) contain several interesting features that challenge previous models.
Several extensions and generalizations of this theory are however necessary for more
comprehensive comparison with experiment. Three immediate extensions that are
necessary are as follows: firstly, the predictions the theory makes for the case of
controlled notch testing need to be developed; secondly, the effect of broader pore
size distributions on the theoretical predictions need to be assessed; finally, the
extension to other loading conditions needs to be developed.
There is also the issue of the reliability distribution appropriate to porous
materials, and this issue will be taken up in a broader context in the fourth part
of this section.

6.6.2 Dielectric breakdown of metal-loaded insulators


The extreme scaling theory of equation (6.27) may also be adapted to the
study of real materials, in this case, insulators containing volume fraction q of small
6. Discussion and experimental implications 221

(nearly spherical) metal inclusions. The identification needed in this case is

V = Ld « (volume of sample)/(average volume of a metal inclusion) (6.65)

Features 1-4 described in the previous subsection then apply equally well to
the dielectric breakdown field of metal loaded dielectrics. A series of experiments
to test the dilute limit singularity as q —* 0 (small volume fractions of metal
inclusions) has recently been published and is presented in figure 6.18.
The experiments were performed on 0.5 mm thick discs of metal loaded
poly-ethylene with 10~5 < q < 5 x 10" 4 . The data of figure 6.18 clearly show the
dilute limit singularity predicted by equation (6.27) and illustrated numerically in
figure 6.6. The unbiased test of figure 6.18b provides strong confirmation of the
analytic form of this singularity (with exponent ß = 1).

6.6.3 A comment on the fractography of brittle materials


The study of fracture surfaces yields a wealth of information about the
effect of microstructure on crack topology. ^ There has been extensive use of this
correlation to understand failure of specific materials systems. A more universal
approach has recently been suggested in which the roughness of fracture surfaces
is characterised by a fractal dimension, Ds. These ideas have been applied to both
ductile m e t a l s ^ and brittle ceramics and glasses^, where a positive correlation
between fracture toughness and Ds has been reported. Although these correlations
are very interesting, we show here that this positive correlation is not a general prin-
cipal, and that without more detailed specification of the microstructure, toughness
and Ds are uncorrelated.
The point is most simply made by considering the two extreme cases
schematically depicted in figure 6.19.
In figure 6.19a, we show a crack propagating in a brittle material contain-
ing random voids of volume fraction, ς^. The roughness of the fracture surface
increases with increasing q± (as occurs for example in figure 6.4), but the fracture
toughness decreases with increasing </χ. For the system in figure 6.19a, there is thus
a negative correlation between Ds and fracture toughness, in figure 6.19b, we show
a crack propagating in a brittle material containing volume fraction, q2 of rigid
second phase inclusions. Again the roughness of the fracture surface increases with
q2, but now the fracture toughness also increases with q2. For the system in figure
6.19b, there is a positive correlation between fracture toughness and Ds. By vary-
ing qi and q2, we anticipate that it is possible to go from a positive to a negative
correlation between fracture toughness and Z)s, so that in general there is no corre-
lation between fracture toughness and the fractal dimension of the fracture surface.
Controlled experiments testing these issues are necessary to quantify the relative
importance of pores as opposed to rigid inclusions on, Ds, fracture toughness and
the correlation between them.
222 P.M. Duxbury / Breakdown

i.oi
I
1
I
I
I
I
It

5c(o)
0.5 k

0.0
2 3
4
q(x10" )
a)

Figure 6.18 a) The dielectric breakdown field of metal-loaded polyethylene as a function of the
volume fraction, #, of metal inclusions and; b) a test of the q —> 0 scaling limit of equation (6.27)
with ß = 1 (from ref. 94).
6. Discussion and experimental implications 223

o 4 /ο ο
o ° / A
o / ο
wO AA cy-
υ L)~vy-l<\.«.i^\
^OvJ

/I
■^ν© Λ / θ Ο /
Vos , T>/ ο°
a) b)

Figure 6.19 Schematic of the cross sections of fractured materials for: a) a material containing
random voids and; b) a material containing random rigid inclusions.

6.6.4 Implications for reliability analysis


Reliability distributions are of vast importance as they are used to extrapo-
late from test loading conditions (where the probability of failure is of order unity)
to design loading conditions (where the probability of failure is typically of order
10~6). Despite the fact that many different types of reliability distribution are pos-
sible, in most materials applications the Weibull distribution is almost exclusively
used. This is partially because it is a very robust distribution that errs on the side
of underestimating reliability, and partially because on the basis of usual reliability
test data it is impossible to differentiate between the Weibull and other reliability
forms. This is clearly seen in figures 6.8a and 6.9 where even with the very large
data set of 2000 samples, it is difficult to distinguish between the Weibull and the
new form proposed in equation (6.39). In real materials testing, far fewer samples
are typically available, with 5 — 10 being usual for industrial component testing and
10 — 200 samples being used in most basic materials research. To predict the design
load for say one part per million reliability, we must extrapolate using parameters
found using the available data sets. In the case depicted in figure 6.8a, such an
extrapolation using first the Weibull and then the modified form led to results
that differed by 30% at the one part per million level. This difference becomes
larger the higher the required reliability, and makes the choice of a good reliabil-
ity distribution imperative for high reliability applications such as the aerospace,
space or nuclear industries (where reliabilities of order 10~8 are required for critical
components).
Since it is very difficult to distinguish between reliability distributions based
on failure statistics alone, there is considerable motivation to determine whether we
can predict the reliability distribution based on studies of the material microstruc-
ture and mode of failure. The idealised models of sections 6.3 and 6.4 make several
predictions in this regard, and three examples of this are:
224 P.M. Duxbury / Breakdown

1. In loosely wound yarns (and many other equal load sharing systems), the relia-
bility distribution is normal in formJ6^
2. In fiber reinforced laminates (with local load sharing) failing by mode I brittle
failure, the reliablity distribution is Weibull with an index that depends on the
width of the laminate (size dependent Weibull modulus).^
3. In brittle porous materials with random near spherical pores, the reliability
distribution is of a modified Gumbel form [38>40'561.
The result 3 is different from the usual Gumbel form which arises in weakest
link arguments. Since the failure in the models of section 6.3 fail at points carrying
the largest loads, the extreme distribution that applies is a maximum extremal
form rather than the minimum extremal form that applies if the material fails at
weak links. In most materials, it is a combination of load hotspots and weak links
that dominate, and more modelling is necessary to determine which of these effects
dominates for a given microstructureJ61^
Although the results 1 — 3 are for idealized models, they show that in
principal it is possible to find correlations between microstructure, damage evo-
lution and the form of the reliability distribution. This sort of information is not
currently employed in reliability studies of material failure and is one area where
more research, both theoretical and using idealized experimental systems, is neces-
sary. This development is critical for the use of new materials in components which
require high reliability.

6.7 Conclusion
In this chapter we have discussed the strength properties of some simple
disordered networks. Using model microstructures, bond constitutive relationships
and damage evolution algorithms, it is relatively straightforward, although CPU
intensive, to numerically generate the complex crack patterns that apply to these
models (see section 6.3.1). These network computational methods are similar in
spirit to the finite element methods often used in fracture analysis ^105\ although
there have been few calculations of breakdown in random media using these meth-
ods.
Network models have predictive power, if trends as a function of disor-
der parameters and systems size can be quantified. Examples of this predictive
power are the analytic forms for brittle strength as a function of porosity given
in equations (6.52)-(6.54). Two experimental consequences of these predictions
are discussed in section 6.6. It is also possible to derive relationships between mi-
crostructure and reliability distributions, and this is illustrated in section 6.3.3.
One prediction of this analysis is that the reliability distribution for porous sys-
tems may be of a modified Gumbel form (see equations (6.33) and (6.39)) rather
than the Weibull form usually used in analysis of strength statistics.
The analytic results of this chapter were derived using a novel combina-
tion of ideas from statistical physics, extreme statistics and fracture mechanics,
References 225

and there is considerable scope for the further development of this combination of
techniques. Using these analytic methods in combination with computer simula-
tions of damage evolution, it should be possible to address a variety of problems in
the fracture of disordered materials. A prelimary discussion of two such problems:
toughening of disordered systems and fractography of random materials is given
in sections 6.4 and 6.6.3 respectively.
As a closing comment, it is interesting that the theoretical discussion of
this chapter, and indeed this book, has been very broad in that electrical, dielectric
and mechanical failure have been studied in parallel. This unified framework for
the study of breakdown phenomena has been lacking in past years and is likely to
be fruitful in the coming years.
I would like to thank my graduate student Yongsheng Li, and collabo-
rators Paul Leath and Paul Beale who have contributed a great deal to the re-
search described here. I have also profited from discussions with Michael Stephen,
Michael Aizenman, Sidney Redner, David Clarke, Robert Cook, Karl Sieradski,
Chris Lobb, Mike Thorpe, Hans Herrmann and Leigh Phoenix. Special thanks to
Stephane Roux and Hans Herrmann for a critical reading of the manuscript. Fi-
nancial support by the Petroleum Research Fund administered by the American
Chemical Society, and the Composite Materials and Structures center at MSU is
gratefully acknowledged.

References
1. G.K. Batchelor, Ann. Rev. Fluid. Mech. 6, 227 (1974); S. Kirkpatrick, Rev.
Mod. Phys. 45, 574 (1971); D. Stauffer, Introduction to Percolation The-
ory (Taylor and Francis, London 1985); Random Media and Composites Pro-
ceedings of the SIAM conference, R.V. Cohn and G.W. Milton eds., (SIAM,
Philadelphia 1989)
2. B.J. Last and D.J. Thouless, Phys. Rev. Lett. 27, 1719 (1971); S. Feng, B.L
Halperin and P.N. Sen, Phys. Rev. B35, 197 (1987); C.J. Lobb and M.G.
Forester, Phys. Rev. B35, 1899 (1987)
3. Y. Kantor and D.J. Bergman, Phys. Rev. Lett. 53, 511 (1984); L.Benguigui,
Phys. Rev. Lett. 53, 2028 (1984)
4. D.R. Clarke and K.T. Faber, J. Phys. Chem. Sol. 48, 1115 (1987)
5. Y.S. Li and P.M. Duxbury, Phys. Rev. B40, 4889 (1989)
6. H.E. Daniels, Proc. Roy. Soc. A183, 405 (1945)
7. B.D. Coleman, J. Appl. Phys. 29, 968 (1958)
8. B.D. Coleman, J. Mech. Phys. Sol., 7, 60 (1958)
9. C. Zweben and B.W. Rosen, J. Mech. Phys. Sol. 18, 189 (1970)
10. D.G. Harlow and S.L. Phoenix, J. Comp. Mats. 12, 195 (1978)
11. S.L. Phoenix, Adv. Appl. Prob. 11, 153 (1979)
12. R.L. Smith, Proc. Roy. Soc. A372, 539 (1980)
13. R.L. Smith and S.L. Phoenix, J. Appl. Mech. 48, 75 (1981)
226 P.M. Duxbury / Breakdown

14. D.G. Harlow and S.L. Phoenix, Int. J. Fract. 17, 601 (1981)
15. D.G. Harlow and S.L. Phoenix, Adv. Appl. Prob. 14, 68 (1982)
16. R.L. Smith, Ann. Prob. 10, 137 (1982)
17. R.L. Smith, S.L. Phoenix, M.R. Greenfield, R.B. Henstenburg and R.E. Pitt,
Proc. Roy. Soc. A388, 353 (1983)
18. S.L. Phoenix and L.J. Tierney, Eng. Fract. Mech. 18, 193 (1983)
19. L.N. McCartney and R.L. Smith, J. Appl. Mech. 105, 601 (1983)
20. R.L. Smith, Adv. Appl. Prob. 15, 304 (1983)
21. R.L. Smith and H.W. Taylor, Operations. Research, 32, 649 (1984)
22. D.G. Harlow, R.L. Smith and H.M. Taylor, J. Appl. Prob. 20, 358 (1983)
23. S.B. Batdorf and R. Ghaffarian, Int. J. Fract. 26, 113 (1984)
24. D.L. Turcotte, R.F. Smalley and S.A. Solla, Nature 313, 671 (1985)
25. R.F. Smalley, D.L. Turcotte and S.A. Solla, J. Geo. Res. 90, 1894 (1985)
26. J. Lomnitz-Adler, Geophys. J. R. Astr. Soc. 83, 435 (1985)
27. J. Lomnitz-Adler, Tectonophysics 120, 133 (1985)
28. L. de Arcangelis, S. Redner and H.J. Herrmann, J. Physique Lett. 46, L585
(1985)
29. P. Ray and B.K. Chakrabarti, J. Phys. C18, L185 (1985)
30. P. Ray and B.K. Chakrabarti, Solid State Comm. 53, 477 (1985)
31. K. Sieradski, J. Phys. C18, L855 (1985)
32. M. Sahimi and J.D. Goddard, Phys. Rev. B33, 7848 (1986)
33. R.B. Stinchcombe, P.M. Duxbury and P. Shukla, J. Phys. A19, 3903 (1986)
34. P.M. Duxbury, P.D. Beale and P.L. Leath, Phys. Rev. Lett. 57, 1052 (1986)
35. L. Benguigui, Ann. Israel. Phys. Soc. 8, 288 (1986)
36. K. Sieradski and R. Li, Phys. Rev. Lett. 56, 2509 (1986)
37. T.L. Chelidze, Pageoph. 124, 731 (1986)
38. P.M. Duxbury, P.L. Leath and P.D. Beale, Phys. Rev. B36, 367 (1987)
39. C.C. Kuo and S.L. Phoenix, J. Appl. Prob. 24, 123 (1987)
40. P.M. Duxbury and P.L. Leath, J. Phys. A20, L411 (1987)
41. D.R. Bowman and D. Stroud, (unpublished)
42. B. Kahng, G.G. Batrouni and S. Redner, J. Phys. A20, L827 (1987)
43. M. Soderburg, Phys. Rev. B35, 352 (1987)
44. A. Gilabert, C. Vanneste, D. Sornette and E. Guyon, J. Physique 48, 763
(1987)
45. L. Benguigui, P. Ron and D.J. Bergman, J. Physique 48, 1547 (1987)
46. E. Guyon, S. Roux and D.J. Bergman, J. Physique 48, 903 (1987)
47. C.J. Lobb, P.M. Hui and D. Stroud, Phys. Rev. B36, 1956 (1987)
48. M. D. Stephens and M. Sahimi, Phys. Rev. B36, 8656 (1987)
49. J. Machta and R.A. Guyer, Phys. Rev. B36, 2142 (1987)
50. Y.S. Li and P.M. Duxbury, Phys. Rev. B36, 5411 (1987)
51. Y.S. Li and P.M. Duxbury, Phys. Rev. B38, 9257 (1988)
52. P.D. Beale and P.M. Duxbury, Phys. Rev. B37, 2785 (1988)
53. P.D. Beale and D.J. Srolovitz, Phys. Rev. B37, 5500 (1988)
References 227

54. D.J. Srolovitz and P.D. Beale, J. Am. Ceram Soc. 71, 362 (1988)
55. M. Octavio, A. Octavio, J. Aponte, R. Medina and C. J. Lobb, Phys. Rev. B37,
9292 (1988)
56. P.M. Duxbury and Y.S. Li, in the Proceedings of the SIAM conference on
Random Media and Composites, R.V. Cohn and G.W. Milton eds., (SIAM,
Philidelphia 1989)
57. B.K. Chakrabarti, Reviews of Solid State Science 2, 559 (1988)
58. D. Sornette, J. Physique 49, 1365 (1988)
59. D. Sornette, J. Physique 49, 889 (1988)
60. B. Kahng, G.G. Bartrouni, S. Redner, L. de Arcangelis and H.J. Herrmann,
Phys. Rev. B37, 7625 (1988)
61. S. Roux, A. Hansen, H.J. Herrmann and E. Guyon, J. Stat. Phys. 52, 237
(1988)
62. H.J.Herrmann, A. Hansen and S. Roux, Phys. Rev. B39, 637 (1989)
63. A. Hansen, S. Roux and H.J. Herrmann, J. Physique 50, 733 (1989)
64. L. Benguigui, Phys. Rev. B38, 7211 (1988)
65. S.K. Chan, J. Machta and R.A. Guyer, Phys. Rev. B39, 9236 (1989)
66. D. Sornette, J. Physique 50, 745 (1989)
67. P.L. Leath and W. Tang, Phys. Rev. B39, 6485 (1989)
68. W. Xia and P.L. Leath, Phys. Rev. Lett. 63, 1428 (1989)
69. W. Newman and A.M. Gabrielov, preprint; A.M. Gabrielov and W.I. Newman,
preprint
70. E.J. Garboczi, Phys. Rev. B37, 9005 (1988)
71. R. Cook, IBM preprint
72. B. Kahng J. Phys. A in press
73. C. Vanneste, A. Gilabert and D. Sornette, preprint
74. W. Weibull, J. Appl. Phys. 18, 293 (1951)
75. E.J. Gumbel, Statistics of Extremes (Columbia University Press, NY 1958); J.
Galambos, The Asymptotic Theory of Extreme Order Statistics (John Wiley
and Sons, NY 1978); M.R. Leadbetter G. Lindgren and H. Rootzen, Extremes
and Related Properties of Random Sequences and Processes (Springer-Verlag,
NY 1983); E. Castillo, Extreme Value Theory in Engineering (Academic Press,
NY 1988)
76. See e.g. A. Kelly and N.H. MacMillan, Strong Solids, 2nd edition (Clarendon
Press, Oxford 1986)
77. D.J. Bergman, Phys. Rev. B39, 4598 (1989)
78. S. Feng, M.F. Thorpe and E. Garboczi, Phys. Rev. B 3 1 , 276 (1985)
79. R.F. Cook and D.R. Clarke, Acta metall. 36, 555 (1988)
80. L. Niemeyer, L. Pietronero and H.J. Wiesmann, Phys. Rev. Lett. 52, 1033
(1984)
81. M. Murat, Phys. Rev. B32, 8420 (1985)
82. H.J. Wiesmann and H.R. Zeller, J. Appl. Phys. 60, 1770 (1986)
83. L. Pietronero, A. Erzan and C. Evertsz, Physica A151, 207 (1988)
228 P.M. Duxbury / Breakdown

84. E. Louis and F. Guinea, Europhys. Lett. 3, 871 (1987)


85. L. Fernandez, F. Guinea and E. Louis, J. Phys. A21, L301 (1988)
86. P. Meakin, Crystal Properties and Preparation 17-18, 1 (1988)
87. E. Hinrichsen, A. Hansen and S. Roux, Europhys. Lett. 8, 1 (1989); A. Hansen,
E.L. Hinrichsen and S. Roux, J. Phys. A 22, L795 (1989)
88. P. Meakin, G. Li, L.M. Sander, E. Louis and F. Guinea, J. Phys. A in press
89. Y.S. Li and P.M. Duxbury, preprint
90. Y.W. Mai and B.R. Lawn, J. Am. Ceram. Soc. 70, 289 (1987)
91. R.F. Cook, C.J. Fairbanks, B.R. Lawn and Y.W. Mai, J. Mater. Res. 2, 345
(1987)
92. K. Kendall, AIP Conf. Proc. A183, 405 (1984)
93. see article by D.P.H. Hasselman and R.M. Fulrath in Ceramic Microstructures:
Their Analysis, Signincance and Production R.M. Fulrath and J.A. Pask eds.,
(Wiley, NY 1966)
94. R.W. Coppard, L.A. Dissado, S.M. Rowland and R. Rakowski, J. Phys. Cond.
Mat. 1, 3041 (1989)
95. Fractography - Microscopic Cracking Processes, C D . Beachem and W.R.
Warke eds. (ASTM tech. pub. 600, 1975); Fractography in Failure Analysis,
B.M. Strauss and W.H. Cullen eds. (ASTM tech. pub. 645, 1977); Fractogra-
phy in Materials Science, L.N. Gilbertson and R.D. Zipp eds. (ASTM tech.
pub. 733, 1979); Fractography of Ceramic and Metal Failures, J.J. Mecholsky
and S.R. Powell eds. (ASTM tech. pub. 827, 1982)
96. B. Mandelbrot, D.E. Passoja and A.J. Paullay, Nature 308, 721 (1984)
97. J.J. Mecholsky, D.E. Passoja and K.S. Feinberg-Ringel, J. Am. Ceram. Soc.
72, 60 (1989)
98. D.B. Marshall and J.E. Ritter, Ceram. Bull. 66, 309 (1987)
99. A. de S. Jayatilaka and K. Trustum, J. Mat. Sei. 12, 1426 (1977)
100. U. Lindborg, Acta Met. 17, 521 (1969); R.A. Hunt, Acta. Met. 26, 1443
(1978); R.A. Hunt and L.N. McCartney, Int. J. Fract. 15, 365 (1979)
101. B.Bergman, J. Mat. Sei. Lett. 4, 1143 (1985)
102. L.A. Dissado, J.C. Fothergill, S.V. Wolfe and R.M. Hill, IEEE Trans, on El.
Ins. EI-19, 227 (1984)
103. Statistical Research on Fatigue and Fracture, T. Tanaka, S. Nishijima and M.
Ichikawa eds., Curr. Jap. Mat. Res. 2, (Elsevier, London 1987)
104. H.L. Harter, Int. Stat. Rev. 46, 279 (1978)
105. A. Needleman, in Theoretical and Applied Mechanics P. Germain, M. Piau
and D. Caillerie eds, (Elsevier, IUTAM 1989)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
© Elsevier Science Publishers B.V. (North-Holland), 1990 229

7. Randomness in breaking thresholds

Lucilla de Arcangelis*

7.1 Introduction
Fracture patterns can appear in a variety of existing materials with shape
and extension that crucially depend on the nature of the medium. Very different
cracks can indeed be observed experimentally in a composite, a fibrous, a porous
or a granular material.^ Moreover, these patterns depend on how the external
stress is applied to the system: if at a point, on a uniaxial boundary or anywhere
else. In fact, not only the final cracks have a different geometry but also their way
to propagate through the system is very sensitive to the distribution of external
stress.
A major effort has been undertaken in the last years in an attempt of
providing satisfactory models^ for the fracture of real, disordered materials (see
chapter 5). One of the most successful approaches simplifies the description of
a continuous material by considering instead a lattice of bonds. Each bond is
described by its elastic and breaking characteristics which are supposed to represent
the system at a mesoscopic level. The choice of the single bond characteristics
together with the breaking rule controlling the failure of a single element of the
system provides a variety of possible statistical models to mimic more or less closely
the fracture of real materials.
The essential ingredient, however, for any realistic model is the presence
of disorder. This can indeed be introduced in different ways: either by choosing a
probabilistic breaking rule or by having a given quenched disorder built into the
system but assigning a deterministic rule for breaking.
In the first type of approach, one can imagine for instance to assign to each
bond in the lattice a probability of breaking proportional to the force applied on
itJ3,4^ By drawing random numbers one then breaks one element after the other
according to the distribution of local probability in a very similar way to the DLA
problem (see chapter 8).
* Service de Physique Theorique, CEN Saclay, F-91191 Gif-sur-Yvette Cedex, France.
230 L. de Arcangelis / Randomness in breaking thresholds

An alternative solution would consist in calculating the elastic energy E


stored in each bond, given the stress applied to it. The probability of breaking is
then defined via the Boltzmann factor e~(3E assigned to each element of the lattice,
where the parameter ß plays the role of inverse temperature J5,6^ This choice would
correspond to a thermally activated breaking process (see chapter 9), appropriate
for instance for a surface of wet sand or clay drying in the sun.
Disorder can be introduced as well by choosing a deterministic breaking
rule, in which case a bond is usually supposed to be ideally fragile with its behaviour
determined by the microscopic breaking characteristics. In the most common case
these characteristics are chosen to have a linear elastic dependence between force
and displacement up to a threshold displacement 6C, beyond which the single bond
breaks irreversibly. Of course, more complicated functions could be considered for
the microscopic characteristics, including for instance fatigue or plasticity effects.
We will see, however, that even this simple choice of the bond being ideally fragile
will give rise to a rich behaviour for the breaking characteristics of the whole
system.
The quenched disorder can now be introduced in several ways. By dilu-
tion, f7_11l that is by setting the elastic constants at random to zero for a given
fraction 1 — p of bonds at the beginning of the process, whereas the remaining
fraction p has the same characteristics with, let us say, a unit elastic constant. The
breaking process can then take place provided that the probability p is greater than
the (elastic) percolation threshold, in order to ensure a non-zero Young's modulus
for the system (see chapter 4).
While this percolation-type of disorder can be appropriate to describe the
breaking of materials such as porous media, the general situation where all the
bonds are present in a disordered network is often more realistic to describe break-
ing processes in disordered solids. To account for this, the lattice can be chosen
initially with all bonds present and the quenched disorder is then built into the
system by assigning to each bond the same breaking threshold but a different elas-
tic constant drawn at random from a continuous distributionJ 12 ' 13 ! The breaking
criterion, whether in the local stress, in the elongation or in the stored energy,
needs to be chosen next to fully specify the model. The dilution disorder can then
be considered as a particular case in which the elastic constants are simply set
equal to zero or one according to a bimodal distribution. Alternatively, the elastic
constants can be all set equal to unity whereas the breaking thresholds 6C are as-
signed at random according to a given distributionJ 13-21 ^ In both cases the disorder
is quenched in the system, that is the values of the elastic constants and breaking
thresholds are set at the beginning and kept the same throughout the process.
In order to assign the random variables, different distributions have been
considered. The most commonly used is the power-law distribution P(X) — (1 —
α)Χ~α with a < 1 and 0 < X < 1, which contains for a = 0 the uniform
distribution. The disorder can be tuned by varying the exponent a, increasing the
width of the distribution for higher values of a.
1. Introduction 231

Furthermore, the Weibull distribution has also been considered, namely


P(X) = mXm-1 e -W*o) m /* 0 m with X > 0, which is empirically found to fit the
distribution of breakdown stress in real materials. The parameter m controls the
degree of disorder in the distribution and it is usually chosen in the range 2 < m <
10, experimentally found to describe a variety of materials. By increasing the value
of m, the distribution exhibits a faster decay, decreasing the quenched disorder
introduced in the system. A wider distribution is instead obtained for small values
of ra. Finally, the constant X0 is an irrelevant parameter that fixes the average
value of X.
The next step toward the definition of a breaking model is the choice of
the physical nature of each bond in the lattice and subsequently of its microscopic
breaking rule. The first model one can imagine is a random network of electrical
fuses,^ which has received much attention due to its intuitive simplicity and ef-
ficient implementation on the computer (see chapter 6 for a detailed discussion
in the case of dilution disorder). Each element of the lattice is an ideally fragile
electrical fuse, that has linear current-volt age characteristics up to a threshold,
where it irreversibly becomes an insulator. By applying an external voltage (or
current) between two opposite sides of the lattice and keeping periodical bound-
ary conditions in the transverse direction, each time step one or several fuses are
burned for which the current passing through exceeds the threshold value, until the
failure of the network is attained and no more current passes through the system.
The analogy between Ohm's law and Hooke's law (current vs. force, voltage vs.
displacement) can be used to simplify the breaking of an elastic medium into a
scalar model. Each site of the lattice is in fact characterized by a single variable,
let us say the voltage, whose values are obtained by solving the Laplace equation
Αφ = 0 for the electrical potential at each site of the lattice, under the assigned
boundary conditions.
This random fuse network model is indeed closely related to the well-known
Dielectric Breakdown Model (DBM) J22^ Here a dielectric medium is made of resis-
tors which can become superconducting bonds with a probability proportional to
the gradient of the electrical potential applied (see also chapter 8). The breaking
process gives rise to a pattern of zero resistance passing through the system. On
the square lattice, which is self-dual, the random fuse network is precisely the dual
problem^23! of the DBM, by simply replacing the voltage drop by the current and
the conductance by the resistance. Therefore, it is likely to expect that, even if
the random fuse network is a rather simplifying model, it will exhibit some of the
more complicated properties of DBM, such as screening and non-local effects. The
essential difference, however, between these two models is that the DBM generates
a connected pattern, whereas in the random fuse model the bonds can be bro-
ken without the connectivity constraint (for relations between DBM and breaking
models see chapter 8).
To better take into account the vectorial nature of fracture of an elastic
medium, more variables than just the electrical potential should be assigned to
232 L. de Arcangelis / Randomness in breaking thresholds

J I II IL

]Qac
ΠΙ II ΙΓ
(a) (b!

Figure 7.1 Schematic representation of the beam model: (a) rotation of one site, (b)flexionof
a beam due to the angles at its extremities (from ref. 18).

each site of the lattice. To this extent, the simplest model one can introduce is
a network of Hookean springs, with linear force-displacement characteristics up
to a breaking threshold. These springs can freely rotate around the nodes of the
lattice and they will break under exceeding elongation or compression. This is the
central-force m o d e l , ^ so called because the force can only act along the axis of
the bond with no bending effects taken into account and it has been discussed in
chapters 3 and 4 in the context of transport properties of elastic systems. Each
site is now characterized by the coordinates of the displacement; for instance, two
variables are needed in the two-dimensional problem. In both these models only
one quantity, however, determines the breaking of a bond, namely an excessive
current through a fuse or an excessive elongation of a spring. Therefore only one
breaking mode (mode I) is possible in this context.
It is known that elastic materials can fail not only by stretching but also
by bending. Therefore, one needs to take these effects correctly into account. The
beam model^25^ consists of a network of beams imagined to be made of some elastic
material with a certain thickness and shear elasticity. They can therefore not only
stretch but also bend under external stress or shear. The implementation of this
model on a lattice can be shown (see chapter 3) to be equivalent to the solution of
the Cosserat equationJ26^ To each site of the lattice one then assigns as variables
the coordinates of the displacement and the bending angle of the beam with respect
to the rest position. In two dimensions, for instance, three variables characterize
each site, namely #, y and the angle Θ (fig. 7.1). Next, the elastic behaviour of
a beam is given by the following set of linear relations^ for the force / and the
shear s acting on it and the moment rrii at site i:
f = a{xi - xj), (7.1a)
s = b(yi-yj + \W-ej)), (7.1b)

na = cl{9i - θά) + - U - Vj + \Wi + l-j ) , (7.1c)


1. Introduction 233

where / is the length of the beam and a, b and c are parameters depending on the
material's Young and shear moduli, the cross sectional area and the moment of
inertia of the beam.
To complete the definition of the model, one needs to define the breaking
rule for a single beam failure. To this extent, a possible choice is to exploit the clas-
sical material science criterion for yielding due to von Mises, discussed in chapter
3. This implies that a beam between sites i and j breaks if
Λ2 + ^ ( Κ · | , Κ | ) ^

where tf and tm are the two thresholds assigned to each beam for the maximum
force and moment applied. Since two mechanisms contribute to the breaking of
a beam, elongation and flexion, two thresholds are needed in this context and
the ratio between the first and the second term in the left-hand side of eq. (7.2),
quantifies the relative strength of two types of breaking of a single beam. This
model is now a fully vectorial model and it can be used to describe the behaviour
of real solids. However, as the random fuse and the central force models, it is based
only on the linear response of a bond (eq. (7.1)). Non-linear effects are in fact not
taken into account, an approximation valid for small local strains, and the breaking
of a beam is supposed to be a sudden, irreversible process.
Finally, to fully specify the breaking model, in addition to the nature of a
bond and its breaking rule one must also define the kinetics of the failure process.
Commonly, it is assumed that the failure of a single bond is a slow process, that
is the time required to break an overstressed bond is much larger than the time
for the stress to reequilibrate throughout the network. Moreover, the time needed
to break a single bond is usually supposed to be shorter the more this bond is
overstressed, so that only- the most overstressed bond is broken at every stage of
the process. Of course, different assumptions are possible, as for instance allowing
all the overstressed bonds to break at once. This choice would instead correspond to
the physical situation that the time for the stress to relax throughout the network
is much longer that the breaking time of any overstressed bond.
All three models presented have received widespread attention in the at-
tempt of finding scaling laws in the fracture problem. In particular, the random
fuse network has been intensively studied in the context of dilution disorder both
numerically and analytically.^8-10! Some of these results, concerning for instance
the size dependence of the breakdown voltage or the distribution of breakdown
strength, are presented in chapter 6.
In the following the interest is mainly focused on the results obtained in
the presence of quenched disorder for the breaking thresholds. In particular, sec-
tion 7.2 will present some analytical results obtained for the random fuse network
with uniformly distributed breaking thresholds, whereas sections 7.3 and 7.4 are
dedicated to recent numerical studies of the three models presented investigating
the existence of universal scaling laws in fracture.
234 L. de Arcangelis / Randomness in breaking thresholds

7.2 Some analytic results

The formulation of scaling laws for a system undergoing fracture is the


major aim of this chapter. Scaling ideas are in fact a powerful tool allowing to
capture the description of the critical behaviour of a system as a simple function
of few variables, like its size or density. Their application in the present case could
therefore lead to the discovery how quantities like the breakdown stress or the
number of bonds broken just before failure depend on the size of the system.
When an external stress (or shear) is applied to a lattice of bonds initially
homogeneous and, let us say, with quenched disorder in the thresholds, one can de-
tect two regimes in the behaviour of the system. Initially one must steadily increase
the external stress in order to break one bond after the other in an uncorrelated or-
der and randomly in space. This regime, in which mostly the weakest bonds in the
system fail, is called the disorder controlled or disordered regime. After a certain
number, n c , of bonds have failed, the correlations between cracks start to domi-
nate the process, small cracks grow and coalesce until a large macroscopic crack is
created to attain the system failure. In this regime the breaking process proceeds
without any further increase in the external stress, led by the local enhancements
in the stress at the tips of the cracks. For this reason this is usually called the
catastrophic regime. The largest external stress, σ0, is applied at the onset of the
catastrophic regime, after a number nc of bonds have failed. These two quanti-
ties are indeed of major technological interest and they can be easily measured
experimentally by imposing an external strain to the system. In the following are
presented some analytical arguments^14! which provide an upper and lower bound
to the behaviour of nc with system size. These approximations are however quite
rough and their validity will be successively verified via computer simulations.
For sake of simplicity, one can consider the simplest model, the random
fuse network on a square lattice, with randomly distributed breaking thresholds.
These are to be uniformly assigned between v„ = 1 — w/2 and v+ = 1 + iu/2, where
w is the width of the uniform distribution. All fuses have unit conductance and
the kinetics of the process is such that the external voltage is adjusted in order to
burn only the most overstressed fuse at each time step.
By increasing the parameter w, i.e. enhancing the quenched disorder in the
system, the breakdown of the lattice can become more gradual and the failure is
led by successive increases in the external potential. Since the lower the value of
w the sooner the catastrophic regime will set in, it is possible that a minimum
value of w, w0, is required in order for the disorder controlled regime to exist. For
w < WQ, therefore the fracture process consists only of a trivial regime, typical of
a perfectly ordered system, in which the removing of the first fuse creates a dipole
field in the network that immediately leads to the creation of a macroscopic linear
crack without any further increase in the external voltage.
It is possible to give an upper bound for the value of w0. Once the first
weakest fuse is removed, the dipole field centered at the position of the missing
2. Some analytic results 235

bond enhances the voltage drop across the bonds horizontally adjacent on the
square lattice to the initial crack. This enhancement factor is weakly dependent
on the position of the initial crack within the finite system and it is found^ to be
equal to a = 4/π on a square lattice in the thermodynamic limit.
For a sufficiently large system of linear size L, the first weakest bond breaks
when the voltage drop across is t;_ = 1 — w/2. The voltage drop across the bonds
immediately adjacent will then be av-. If this value exceeds the breaking threshold
of the strongest possible fuse in the system, namely v+, then the initial bond failure
will necessarily lead to the formation of a linear crack breaking the system apart.
Imposing therefore the condition v+ — av_, one obtains

w0 < 2 (^r1^-) « 0-2404 (7.3)


\ 4 + 7Γ/
as the critical value for the initial single bond crack to be unstable to further
breaking on a square lattice.
Given a system of linear size L and a distribution of width w, an interesting
question is how the breakdown strengths of successive breakings are distributed.
If an external voltage per unit length v is applied, then the probability that a fuse
burns at this voltage drop is p — (v — v-)/w. Therefore, if all the L2 vertical bonds
have the same voltage drop across, the probability that none of them breaks is

V(v) = (l-p)L\ (7.4)

One can then calculate the average voltage per unit length at which the weakest
bond will break, i.e. (υ^) = /J+ vV\v)dv, where V'(v)dv represents the probability
that the weakest bond breaks when the potential is between v and v + dv. By
means of eq. (7.4) and integrating by parts one obtains

{Vh) = v + (7 5)
- Ζ^ΤΪ ·
since the strengths of the L2 vertical bonds are uniformly distributed over the range
w, with the typical difference in strength between two successive breakings of the
order w/L2.
Suppose now that a sequence of n weakest bonds has failed, creating n
isolated single-bond cracks spatially uncorrelated in the system. From eq. (7.5) it
follows that the breaking strength of the n-th of such weakest bonds goes as

(vw(n)> ~ v- + nw/L2 (7.6)

so that the enhanced voltage drop at the tip of each crack is a(vw(n)). Moreover,
once each single bond is broken, the number of bonds at the edge of a crack, which
sustain most of the enhancement in the voltage, increases by two on the square
lattice. By assuming now that these 2n edge bonds are statistically independent,
236 L. de Arcangelis / Randomness in breaking thresholds

from eq. (7.5) it follows that the average strength required to break the weakest
edge bond is given by
M n ) ) ( 7 7 )
= V- + 2^Tl- ·
In order to have that the next bond to be removed is one of the edge bonds, so
creating a crack of size two, the equality αζυ^η)) = (ve(n)) must be satisfied. This
gives the equation
<7 8>
«-«•--(ϊίϊ-^)· ·
for the number n of bonds to be cut in order to grow a crack of size two.
The total number of bonds to be removed in order to reach the catastrophic
regime and therefore the final failure can also be obtained from eq. (7.8) by means
of the so-called unstable-crack approximation}1^ If one assumes in fact that a crack
of length greater than one is immediately unstable to further breaking, the solution
to eq. (7.8) provides a lower bound for the number nc of bonds to be cut to reach
catastrophic failure. Such solution strongly depends on the width of the breaking
threshold distribution. In particular, if w < 2, that is if V- ^ 0 and the distribution
has a finite cutoff, the solution for n c gives a finite number independent of the
system size, i.e. nc ~ L°. This result is independent of the length of the initial stable
crack and can be generalized to cracks of any length providedg the assumption of
spatial independence in the initial sequence of cracks holds.
If instead the distribution has no lower cutoff, i.e. v_ = 0 or w = 2, then
the unstable crack approximation for eq. (7.8) gives as lower bound nc ~ L/y/2a
for the number of broken bonds at the onset of catastrophic failure. This result
constitutes a lower bound for n c , since it strongly underestimates the number of
bonds cut in the first regime. In fact, it has been observed by numerical simulations
that cracks of small size can indeed grow and then stop without attaining the final
failure or they can coalesce into larger cracks. These events are neglected in the
previous approximation which, it should be stressed again, is strongly based on the
assumption of spatial independence for the single-bond cracks.
In order to derive other criteria for crack growth, an alternative argument
has been introduced,^ the so-called dilute-crack approximation. This allows cracks
of length greater than unity to be stable to further breaking but assumes that these
cracks are sufficiently diluted in the system so that the interactions among them
can be neglected. Moreover, cracks are assumed to be straight and to grow only at
the tips, advancing by one lattice spacing. Within these assumptions it is possible
to calculate the crack size distribution and the average breakdown strength of
the network. The approach is then very similar to the one applied to the random
fuse network with dilution disorder, where such distribution is an exponential and
the average breakdown strength decays as l / \ / l n L in two dimensions^ (see also
chapter 6).
In the present case, one can express, in absence of correlations, the proba-
bility of breaking the n-th bond in a crack as pn = (vn — V-)/w, where vn = anv
2. Some analytic results 237

is the enhanced voltage drop at the tip of a crack of length n. Asymptotically vn


is found to vary as®
vn ~ v(l + cy/n), (7.9)
where v is the external voltage applied per unit length and c is a constant. The
probability then for a linear crack of length n to exist is simply

P(n) = poPlp2 · · · pn(l - p n + 1 ) 2 , (7.10)

where the assumption has been used that a crack can propagate only at the tips.
By inserting the asymptotic form (7.9) for vn in eq. (7.10), the integral for In P(n)
can be performed for large n. Furthermore, by considering only the behaviour of
In P(n) for voltages at the tip of the crack very close to the highest threshold v+,
one finds that lnP(n) ~ — n. It follows that the total number of cracks of size n in
the system scales as N(n) = L2P(n). This exponential dependence of N(n) on the
size of the crack will further imply® that the distribution of breakdown strengths
has the form of an exponential of an exponential rather than the expected Weibull
distribution, experimentally found in breakdown testings. This last result is also
found in the random fuse network with dilution disorder, where, however, it is
possible that exactly at the percolation point, where N(n) is a power law in n, the
Weibull form is the correct distribution of failure stresses (see chapter 6).
By setting the number N(n) equal to unity for the longest crack in the
system (extreme-value statistics), one finds that the length of such longest crack
goes as niongest ~ InL. The system then becomes unstable when such a crack,
for which vniongest ~ v+-> exists and, by inserting this relation in eq. (7.9), the size
dependence of the breakdown voltage is found, as for the case of dilution disorder,
to be
<«6> ~ -%= (7.H)
V InL
which combined to eq. (7.6) gives

nc ~ ^ (7.12)

for the number of broken bonds at the onset of catastrophic fracture.


This latest result constitutes an upper bound for n c , it is in fact obtained
assuming that the initial cracks are sufficiently diluted in the system to be uncorre-
lated. Taking into account interactions among cracks leads to mutual enhancement
effects, which tend to organize cracks in space bringing the system earlier into the
catastrophic regime. The quantity nc is therefore bounded between the two be-
haviours, n c ~ L and nc ~ L 2 /VT, and numerical simulations presented in the
following sections will shed further light on its scaling properties.
The presented results can be easily generalized to different distributions
of quenched disorder or to higher dimensions. In particular, one can start with
the more general power law distribution, P(v) = (v — v_)^, where V- φ 0 is the
238 L. de Arcangelis / Randomness in breaking thresholds

lower cutoff. It is possible then to follow also in this case the same arguments just
developped to obtain the quantity n c . For instance, eqs. (7.6) and (7.7) generalize
respectively to (vw(n)) ~ V- + bw(nlL?y and (ve(n)) = v_ + ανυ/2ημ, with μ =
l / ( / ? + l ) . Eq. (7.8) analogously follows and therefore the dependence nc ~ L within
the unstable-crack approximation.
Furthermore, the extension to the three-dimensional case, more relevant
experimentally, could be easily performed. In this case, the probability for a single-
bond crack to propagate behaves very similarly to the two-dimensional problem
. The smaller enhancement of the current at the tips of a crack in 3d competes
with the fact that a crack can propagate in more ways. It would be however very
interesting to investigate how typical structures, like planar cracks, observed ex-
perimentally form and evolve on three-dimensional lattices.
An approach analogous to the dilute crack approximation has also been
applied to the random fuse network with dilution disorder (see chapter 6) and
to elastic modelsJ 1 ^ For both electrical and mechanical failure, it is possible to
derive the scaling behaviour for the enhanced current or stress at the tips of an
isolated crack. Analogously, the dependence of the strength of the system with the
square root of the crack length (eq. (7.11)) can be obtained either by field intensity
calculations (see chapter 6) or else within the classical energy balance arguments
due to Griffith (see chapter 5). Furthermore, the same square root dependence is
also found in the dual problem of the random fuse network in two dimensions, the
failure of a dielectric medium. The dielectric problem exhibits, however, a different
behaviour in three dimensions, where the most critical defect has a finger-like (Id)
structure rather than the planar (d — 1) one of the elastic and fuse problems.
Therefore, in this case the square root behaviour of the electric field (or the stress)
at the tips is replaced by a linear dependence on the crack length.
Two adjacent cracks separated by an unfailed region, whose size is small
with respect to the length of the two defects, represent a different problem, where
spatial correlations cannot be neglected. In this case the field intensity calcula-
tions^ lead to a linear, rather than a square root, dependence of the enhanced
current at the tips. This scaling behaviour is again characteristic also for elastic
modelsJ11^ Moreover, in these models the crack size distribution is found to have an
exponential behaviour, whereas the failure stress distribution has an exponential
of an exponential dependence, as in the electric problem, instead of the Weibull
distribution. The study of the critical current in an inhomogeneous superconductor
has also been addressed as a percolation breakdown phenomenonJ27^ In particular,
it has been found that the critical current goes to zero logarithmically with the size
of the system and that the critical current distribution has a double-exponential
form.
Finally, a very recent p a p e r ^ has presented a Flory-type of approach to
describe the early stage of rupture for a network with quenched disorder in the
breaking thresholds. The approach consists of considering the enhanced stress in
the vicinity of an existing crack as an energetic, favourable field for the nucleation of
3. Breaking characteristics 239

the crack. Conversely, the quenched disorder in the threshold strength of each bond
plays the role of an entropic contribution with an effective hindering effect on the
breaking of a neighbouring bond. The rupture of the system is then studied with
arguments similar to the Flory approximation for polymers, since the propagation
or creation of a crack is governed by the competition between the energy-like and
the entropy-like term.
More precisely, for the case of a uniform distribution of thresholds with a
width w < 2, as the external stress is increased from zero one bond is broken after
the other. It is found that, in the early stage of t bonds is cut in a system of a given
size, at a finite typical distance one from the other, giving rise to a diffuse damage
in the neighbourhood of the crack. When w approaches 2, the process is strongly
dominated by the entropic term and more and more bonds are broken. The typical
distance between them goes like the size of the system, whereas the value of the
external stress applied tends to zero as w —> 2. If instead the width w goes to zero,
the rupture is entirely energy-dominated and the system undergoes catastrophic
fracture. All the results obtained within this Flory-type of approach are therefore
in agreement with the predictions of the unstable crack approximation.
The same arguments have also been applied to the case of the power law
distribution of thresholds in the vicinity of the zero rupture stress. The results
obtained strongly depend on the exponent of the distribution of thresholds, which
determines whether the rupture is dominated by the entropic or the energetic term.

7.3 Breaking characteristics


In order to verify the above predictions for the scaling behaviour of the
breakdown voltage and the number of bonds cut at the onset of the catastrophic
regime, extensive numerical simulations have been performedJ 16-19 ] The aim is to
see if scaling laws can be formulated in the fracture problem and to verify their
validity with respect to different parameters, like boundary conditions, nature of
the bonds and distributions of quenched disorder.
The fuse, the central force and the beam model, as defined in the Introduc-
tion, have been studied in two dimensions with unit bond conductance or elastic
constant and quenched disorder in the breaking thresholds. These are assigned ac-
cording to the power law and uniform distribution for all three models and the
Weibull distribution for the fuse network. In all cases the exponents x, for the
power law, and m, for the Weibull distribution, are varied over a wide enough
range to analyse the behaviour of the system in the limit of very small and high
disorder. Moreover, the external stress is applied at two bus bars placed at the top
and bottom of the lattice whereas periodical boundary conditions are kept in the
transverse direction.
Let us specify now the breaking process. In the fuse modelJ19^ once a unit
external voltage is applied, the Laplace equation is numerically solved at the sites
of the square lattice to find the current i passing through each fuse. Then, the
240 L. de Arcangelis / Randomness in breaking thresholds

bond with the largest ratio A = \i\/ic, where ic is the current breaking threshold,
is chosen and irreversibly removed. This procedure is equivalent to setting the
external voltage to the value λ so that only the most overstressed fuse burns at
a current |i| = ic each time step. The process goes on until the whole network
becomes disconnected. Each time a fuse is removed, the local currents for the new
configuration of bonds must be calculated again so that the procedure is quite
time consuming. An efficient way to solve this problem numerically is using the
conjugate gradient methodJ29^ This is done in the present case with an accuracy
e = 10~12 on the value of the local currents.
Analogously, in the central force m o d e l ^ a triangular lattice of Hookean
springs with unit elastic constant is considered. After imposing a unit external
displacement to the rigid bus bars, the distribution of local forces at equilibrium is
calculated numerically with conjugate gradient techniques. As in the scalar case,
once the local forces are determined for a unit external displacement, the bond for
which the ratio λ = | / | / / c is largest will be removed and the process continued
until final failure is attained.
Finally, in the beam m o d e l ^ a unit external strain is applied, as elongation
or shear, and the local forces and momenta are determined according to eq. (7.1).
The breaking of a single bond, zj, proceeds by determining the smallest value of λ
for which (see section 5.4)
fX\2 + max(\mi\\,\mj\\) = χ
tf ) tm
More precisely, this is done by calculating at each site of the square lattice the
coordinates of the displacement, x\ and y^ and the bending angle 0*. From there are
determined the values of / and ra^ to provide the minimum λ which identifies the
beam to be removed. The material constants a, b and c in eq. (7.1) are arbitrarily
chosen equal for each bond and fixed at the beginning and then set to infinity once
a beam is broken. The procedure is carried on, as in the previous models, breaking
one beam at the time and recalculating x^ yi and 0; using the conjugate gradient
method.
As the process goes on, the complete history of the rupture is recorded,
namely the external force F(n\ the external displacement λ(η), etc. are monitored
at the breaking of the n-th bond. Moreover, for the various system sizes L and the
different distributions analysed, these quantities are averaged at each stage of the
rupture over an ensemble of several initial configurations of quenched disorder at
fixed n.
Each time a given bond is broken, the external elongation (voltage) needed
to do so is just given by λ. Therefore the total external force (current) applied to
break such bond is given by F = YX (I = GX), where Y (G) is the elastic modulus
(conductance) of the system.
The force-displacement characteristics F(X) (I(X) for the scalar model)
contains interesting information concerning the rupture process and can be easily
3. Breaking characteristics 241

F
30

1.0

0.3

0.1

0.1 1 10 100
λ

Figure 7.2 Log-log plot of the force F against the displacement λ at the rupture of individual
bonds in the beam model. An external elongation is applied to a system L = 16 with a uniform
distribution (x — 0) and the average is performed over 60 samples (from ref. 18).

measured experimentally. Fig. 7.2 shows an example of such characteristics for the
beam model. The two different regimes previously introduced can be easily detected
here. In the initial stage of the process the external force steadily increases and is
proportional to the displacement λ. This is the disorder controlled regime where
the weakest bonds in the system are broken randomly in space. Several microcracks
are at this stage initiated at widespread locations^13! and the statistical fluctuations
are small. In other words, in the first regime of the process the distribution of local
stresses on the single bonds is still quite uniform over the whole lattice. Therefore,
if the breaking thresholds are distributed widely enough, the breaking of the bonds
is controlled only by the single bond strength and not by the local force applied.
The larger the amount of disorder in the distribution of breaking thresh-
olds, the wider is the region of linear behaviour, until the maximum value for F
is reached, a value representing the breakdown force for the system. After going
through the maximum the catastrophic regime sets in and one can actually apply
each time a smaller force in order to break the next bond. This region has large
statistical fluctuations and is accessible experimentally only by imposing the ex-
ternal strain, since by imposing the external stress all the bonds in the system will
simultaneously break.
Of major interest is the analysis of the system size dependence of the break-
ing characteristics. By appropriately rescaling the axis, one can in fact try to col-
lapse the data from different system sizes onto one curve. For the force-displacement
characteristics this collapse is obtained using the following scaling ansatz:

F = Ιαφ(ΧΣ~β), (7.14)

where φ is a scaling function.


242 L. de Arcangelis / Randomness in breaking thresholds

1 1
1.0 1 1 i 1f
F/r
M fc
yA
0.8 f
k \l-=2
0.6 [- / / \ 6
5C

0.4
] V \ L = /;
- / U \J^= 1 ^ - ^ -
0.2 -/ \L = 3 2 \
—-,^^^-
/ , 1 i 1 i "T 1 1
8
1.0 XL-0.75
\llm
(a) (b)

Figure 7.3 (a) Force-displacement breaking characteristics for the beam model with x = 0.5
and both axis rescaled by L - 3 / 4 for different sizes L. The data have beensmoothened to reduce
statistical fluctuations.^18! The insert shows the characteristics of a single beam (from ref. 16). (b)
Rescaled force-displacement relation with α ~ β ~ 3/4 for the central force model with a uniform
distribution and system sizes L = 4,8,16,24 (from ref. 17). (c) I/La versus V/Lß for the scalar
model and the Weibull distribution with m = 10. The system sizes are L = 4,8,16, 32,64,128
and the average is taken from over several thousand for L up to 16 to 30 samples for L = 128
(from ref. 19).
3. Breaking characteristics 243

The role of this relation is to define some reduced variables, FL~a and
_/3
λΙ/ , which are related through a function independent on the lattice size. Fig. 7.3
shows the rescaled characteristics for the three different models and different dis-
tributions. The collapse works in the first regime for a ~ β ~ 0.75, whereas there is
no evidence for a scaling behaviour in the catastrophic regime. Further attempts to
collapse the data in this second regime not with a simple power law as in eq. (7.14),
but using a more complicated scaling behaviour, such as L a (ln L)@, have also failed
to give a better collapse of the data. In particular, it is interesting to notice that
the data do not follow the scaling with L/\/TnX, predicted by the analytical ap-
proximations (eq. (7.11)) presented in the previous section. Furthermore, the two
exponents a and ß in eq. (7.14) are treated as two independent parameters and
they are found to be equal within error bars of approximately ten percent. This
result, a ~ β ~ 0.75, quite different from the intuitively expected a = d — 1 and
β = 1, seems also to be independent of the model considered and of the distribu-
tion of quenched disorder, provided that it is wide enough so that the first regime
exists.
When the amount of disorder in the system is reduced the first regime
progressively disappears and, after the breaking of few initial bonds, the system
immediately enters the catastrophic regime. This is seen in fig. 7.3c, which shows
the breaking characteristics for the fuse model with a very narrow distribution of
thresholds, the Weibull distribution with m — 10. Here, the breaking of the first
few fuses is already at a value of the current very close to the maximum current
applied to the network. Moreover, in the catastrophic regime the characteristics
bend back toward smaller values of the voltage so the final breakdown voltage is
actually quite lower than the maximum one.
This behaviour is reminiscent of the situation of a crack propagating in
a self-sustained way through a nearly ordered system. In this case, after the first
bond cut the crack propagates without further increase in the external potential,
driven by the local enhancement in the current at its tips. This is the behaviour
that, for instance, would be obtained if the distribution of breaking thresholds is
a delta function and a constant constraint is applied to the system.
Another interesting quantity to investigate is the average number n of
bonds that have been broken at a given stage of the rupture process. The rela-
tion F(n) then monitors the external force (current) applied for a given number of
bonds cut. Again to determine the scaling behaviour of F and n, one can try to
collapse the data for different L's on a single curve by postulating a scaling ansatz.
Fig. 7.4 shows that a scaling of the form
F = Laj{>(nL-^), (7.15)
where ψ is a scaling function, holds in the disorder controlled regime. As from
eq. (7.13), the exponent a is found to be a ~ 0.75, whereas 7 ~ 1.7. Moreover, the
value of the exponents α, β and 7 are found to agree in the three models within
typical error bars of ten percent. If the width of the distribution of quenched
244 L. de Arcangelis / Randomness in breaking thresholds

0.03

0.6 1.2 1.6

0.25

1.4 1.8

Figure 7.4 I/La versus n/L1 for the scalar model with the power-law distribution (x = 0.5) for
L up to 64 (a), and with the Weibull distribution (m = 2) for L up to 128 (b) (from ref. 19).
Data collapse of F/L3^4 versus n/L1·65 for the beam model with a uniform distribution and an
external elongation applied (c) (from ref. 18).
3. Breaking characteristics 245

disorder decreases, the first regime disappears and, as for eq. (7.13), it is not
possible to collapse the data on a single curve.
The limit of infinite disorder has been analysed^15! for the fuse network
with random thresholds. More precisely, once an external voltage is applied, it is
possible to partition the bonds in the network into two classes: the bonds carrying
a current, belonging to the so-called backbone; and the ones carrying no current,
called dangling ends (see also chapter 4). The hypothesis of infinite disorder consists
in choosing at random the bond to break, assigning equal probability to each bond
in the backbone.
It has been shown that the so-defined process is very close to a percola-
tion problem, where instead the bonds are broken completely at random. In the
early stage of breaking the two problems are indeed the same. For this reason, one
refers to the rupture in the infinite disorder limit as a screened-percolation process.
This relation to a percolation phenomenon shows that the breaking process can
be seen as a critical phenomenon, where the width of the distribution controls the
cross-over between the regime dominated by the quenched disorder and the catas-
trophic regime. In particular, it is possible to derive some of the critical exponents
just before failure in the limit of infinite disorder as function of the well-known
percolation exponents (see section 4.6).
As already pointed out, the two exponents a and /?, treated as independent
parameters, turn out to be equal from the data collapse. It is therefore interesting to
discuss some arguments about the linear dependence of F on λ at the beginning of
rupture. For the scalar model this dependence is easily accounted for by considering
that at the early stage of breaking the system still behaves as a macroscopic,
homogeneous resistor. The whole network then obeys Ohm's law, I = G F , where
the conductance scales as G ~ Ld~2, as for a homogeneous system. It follows that in
two dimensions the external current is simply proportional to the external voltage.
Similarly, in the context of elastic models it is possible to formulate an
effective-medium type of a r g u m e n t ^ at the beginning of rupture, which predicts
the observed result that the external elongation λ(η) has a linear dependence on
the number of bonds cut in the early stage of the process. Let us assume that
the strain applied to the lattice is uniformly distributed throughout the system of
size L, as if it was perfectly homogeneous. Then, due to the periodical boundary
conditions, the horizontal bonds do not carry any force, whereas the bonds in the
other two directions of a triangular lattice are equally elongated by the amount
/
ε = ( v 3/2)A/L.
The bonds that break when λ is applied, if simultaneous breaking is as-
sumed, are those bonds whose threshold is smaller than ε. Their number n for a
uniform distribution between zero and one is simply n = 2L2e, which leads to

A(n) = - ^ n / I (7.16)

in very good agreement with the numerical dataJ 17 ! On the other hand, the external
246 L. de Arcangelis / Randomness in breaking thresholds

I r —i
(o)
300 - Fb
(♦)
100 - 10
So ♦

xb 30 ♦ - 3

(Δ) ♦ Δ
10
♦ Δ
- 1

3
Δ
Δ
- 0.3
Δ
I L —J
100

Figure 7.5 Log-log plot of n\> (o), Ab (Δ) and Fb (o) as function of L for the beam model with
a uniform distribution and external elongation applied (from ref. 18).

force applied when the n-th bond is broken is numerically found to follow^

F(n) = λ(η)(1 - kn/L2), (7.17)

where k is a constant dependent on L. The second term on the right-hand side rep-
resents the first linear correction in n to the Young's modulus Y(n) = F(n)/\(n).
Now, the elastic modulus is very close to one at the beginning of rupture, so that
the linear dependence of F(n) on λ(η) follows and therefore α ~ β.
The scaling dependence of the number of bonds broken at the maximum
of the characteristics can also be obtained within this argument. By inserting
eq. (7.16) into (7.17) one gets F{\) = λ(1 - ky/3/LX). Differentiating F{\) to
find its maximum, it follows that Xh = L/(2y/Sk) ~ iß and from eq. (7.16)

rib Lß+\ (7.18)

which is in good agreement with the values of 7 and ß found numerically.


The maximum of the characteristics is indeed a very interesting point to
investigate further, since the stress there represents the breakdown stress for the
network and the number of bonds broken up to that point is precisely the quantity
n c introduced in the previous section, whose scaling behaviour has been predicted
by different approximations. Fig. 7.5 shows the finite size scaling behaviour of the
force F b , the displacement Ab and the number of broken bonds n b at the maximum
of the characteristics. In agreement with the findings from eq. (7.15), nh follows a
power law scaling with an exponent of about 7/4 ~ /?+l, as predicted by eq. (7.18).
This value seems also to be quite independent on the quenched disorder distribution
and on the boundary conditions. This scaling result is instead in contrast with the
predictions of the unstable-crack approximation (see section 7.2) for a distribution
3. Breaking characteristics 247

J I I L..JJ I l I
3 10 30 100 3 10 30 100
L

Figure 7.6 Number of bonds cut when (a) the maximum force is reached and (b) when the
system falls apart, in a log-log plot versus L for the scalar model with x = 0.0 (α), x = 0.5 ( Δ ) ,
m = 2 ( · ) , m = 5 ( x ) and m = 10 (o); and the beam model for an external elongation with
x = 0.0 ( + ) , x = 0.5 ( v ) , x = - 1 . 0 (A) and for an external shear with x = 0.0 (▼). The full
lines are guides to the eye of slope 1.7 (from ref. 16).

with a lower cutoff. There, the quantity rib is predicted to be a finite number
independent on L.
It is also observed that the force F^ and the displacement Ab (fig. 7.5) do
not lie on a straight line and therefore they do not show any simple power law
behaviour. The data have in fact a definite curvature probably due to strong finite
size corrections or logarithmic prefactors. In any case, the effective slope of Ab is
larger than unity and the curvature of the data is positive. This fact excludes a
size dependence of the type L/y/lnL, as predicted by previous approximations
(eq. (7.11)). Finally, it is worth noticing that also the finite size scaling analysis
indicates that F\> and Ab have essentially the same L-dependence, confirming that
the elastic modulus is independent of L within statistical error bars.
Another interesting point in the process is the end of the breaking charac-
teristics, that is the point where the last bond breaks before the system finally falls
apart. This point provides information both about the catastrophic regime and the
overall process. Fig. 7.6 shows the system size dependence of rib, the number of
bonds cut at the maximum, together with rif, the total number of bonds cut during
the whole process. These two quantities exhibit the same scaling behaviour with an
exponent roughly equal to 7/4 and independent on the choice of the distribution.
However, as the width of the distribution of quenched disorder becomes smaller,
the quantity r?b crosses over to a L° dependence, whereas rif tends towards a lin-
ear behaviour with L. This is in agreement with the picture that for decreasing
disorder the first regime disappears (n b —► 0) and the system is broken by one
248 L. de Arcangelis / Randomness in breaking thresholds

linear macroscopic crack. It has also been numerically verified^ that the length
of the spanning crack, causing the system failure, scales with a fractal dimension
D ~ 1.1 ± 0.1 for all distributions considered.
Moreover, the total number of bonds cut scales with an exponent close
to 1.7 which is by coincidence the value of the fractal dimension of the pattern
obtained in the dielectric breakdown (DB) problem (see chapter 8). It has already
been pointed out in the introduction that this is the dual problem of the random
fuse network in two dimensions. It is, therefore, interesting to find that the DB
and the failure pattern are apparently characterized by the same exponent, even if
the fracture process does not generate a connected structure, contrary to the DB
problem.

Exponent Fuse model Central force Beam model


a 0.88 ± 0.05 0.75 ± 0.08 0.78 ± 0.08
ß 0.84 ± 0.04 0.75 ± 0.08 0.77 ± 0.08
7 1.71 ±0.09 1.72 ±0.18 1.75 ±0.18
In nf / In L 1.71 ±0.09 1.6 ± 0 . 2 1.75 ±0.18
Table 7.1 Values of the scaling exponents found in the three different models: a for the external
force, ß for the external elongation, 7 for the number of bonds cut at the maximum of the
characteristics and the exponent for the total number of bonds cut.

Finally, one should stress again that the values for the exponents presented
here have been determined from independent simulations on the three different
models and they have been found to agree within error bars (table 7.1). More-
over, their value is independent of the boundary conditions or the distribution of
quenched disorder, provided that this is wide enough so that the disorder con-
trolled regime exists. For very small disorder these exponents cross over to the
values expected for a homogeneous system.
In addition, the s t u d y ^ of the scaling behaviour of the breakdown strength
on hierarchical lattices with a distribution of breaking thresholds has also confirmed
that the exponent a is independent on the width of the distribution and the type of
hierarchical organization. In the case of hierarchical lattices with a redistribution of
currents on one generation only, the breakdown strength has been found to depend
as F/L oc log(logL) on the lattice size. Moreover, preliminary simulations^31! on
the three-dimensional central force model have proposed a ^ 2.0.
In conclusion, the study of the scaling properties of the breakdown quanti-
ties has given some evidence for universality in this problem. This means that the
scaling laws found in fracture depend neither on the specific details of the models
nor on the microscopic inhomogeneities introduced on the system. This finding can
therefore be considered as a novel approach to studying the failure in real systems.
4. Multifractality 249

104

10 2

10°

io-2

10 6
J i L
0 30 60 90 120
n

Figure 7.7 Logarithm of h% in the beam model against the number n of broken beams for a
uniform distribution and external elongation in a system of size L = 16. The average is taken
over 500 configurations (from ref. 18).

7.4 Multifractality
Not only the macroscopic breaking properties presented so far are poten-
tially of great technological importance. Indeed materials scientists and mechanical
engineers are often interested in local properties, like the distribution of local stress
or strain throughout the system. The study of these local effects is in fact very help-
ful to better understand the mechanism of rupture and can give useful insight in
the construction of new, stronger materials.
In the context of the simple lattice models presented here, the study of
the local properties, namely the current through a fuse, the force on a spring and
the force or shear exerted on a beam, will give further information on the physical
behaviour of the system during the two detected regimes of rupture.
The first quantity that can be analysed is the overstress on each of the
bonds cut, namely the ratio of the force applied on it to its breaking threshold,
as function of the number of broken bonds. In particular, for the beam model two
such quantities can be defined, one for the elongation and one for the flexion of
a beam, corresponding to the first and second term in the criterion (eq. (7.2)).
Fig. 7.7 shows the behaviour of the first quantity, h2 = {f/tf)2, for a given system
size during the rupture process. The range of variation of h2 is over many orders of
magnitude: at the beginning of the process h2 is indeed very large due to the small
values of tf which determine the breakings. This happens in the disorder controlled
regime, where in fact the weakest bonds are broken.
As the process goes on, h2 decreases very slowly and monotonically with n
as a consequence of the monotonic increase in A (fig. 7.2). The analogous plot for
the overstress due to the flexion of the beam shows a very similar behaviour within
250 L. de Arcangelis / Randomness in breaking thresholds

< n ( i )> I- ' ' 9 0 0


<n(i)>
700
800
500
j

:]v..
600 300

- i 100

1,. j - T T- t I ,
(.00
• · ·
1Γ* ir 8 10-12 j
i
1

200 [ • ·
L,..JM 1
IIP 10"' 101-10 IQ" 14 j

Figure 7.8 Average distribution of local current (n(i)) versus lnz for the fuse model, L = 64,
with a uniform distribution (50 configurations) and the Weibull distribution m = 10 in the insert
(100 configurations) just before the system breaks (from ref. 19).

the statistical fluctuations clearly for both quantities. However, the overstress due
to flexion seems to decrease more slowly as the process goes on, so that it is actually
the elongation given by h\ which controls the behaviour of λ as a function of n.
This depends on the value of the ratio between the widths of the distributions of
the thresholds tm and tf (see eq. (5.13) and discussion thereafter).
If in the controlled regime the choice of the bond to be broken is mainly
determined by its threshold value, in the second regime the local stress is the crucial
quantity that leads the process. It is therefore interesting to investigate how this
local stress is distributed in the system at different stages of the process in order
to understand the mechanisms that control the propagation of a crack and lead to
final failure.
In order to do so, the distribution of the force applied on each bond has
been studied at the very end of the catastrophic regime when a single bond still
holds the system together. In the case of the fuse model this quantity is precisely
the distribution of local currents across each fuse in the network, which has been
first introduced in the context of percolation J32'33! There the investigation of the
current distribution and its moments has led to the discovery of the multifractal^
properties of the incipient infinite percolation cluster at the percolation thresh-
old. The general formalism for the multifractal analysis is extensively discussed
in chapter 4, with particular attention to the random resistor network problem.
The methods presented there are quite general, therefore the reader is advised to
address to that chapter for further details, not discussed in the present context.
Fig. 7.8 shows the current distribution (n(i)) for the random fuse network
just before the last fuse is burned for very weak and large disorder. In the first case,
the system has a very narrow distribution of thresholds (Weibull with m = 10) so
that, after breaking rif ~ L fuses, the distribution of local currents is still strongly
4. Multifractality 251

peaked around its most probable value. It is therefore not so different from the
delta function behaviour found before the breaking process has actually started.
As the disorder increases, the distribution spreads over a wider range to-
ward smaller and smaller values of the current. This effect is due to the creation
of regions of very strong enhancement in the current, as more and more bonds are
cut, together with large areas quite well screened and therefore experiencing small
local stress. In this case of high disorder, the distribution then exhibits features
more similar to the voltage (or current) distribution in percolation at the critical
threshold.
Of course, the main interest is to obtain the scaling behaviour of this distri-
bution of local forces with the system size, that is to find out if the quantity (n(f))
has a power law dependence on the system size L with a given critical exponent.
Usually, this is the behaviour detected in critical phenomena as, for instance, for the
cluster size distribution^35^ in percolation. There the whole distribution increases
with a simple power law behaviour as the system size increases: all the regions in
the distribution have the same scaling dependence and the physical behaviour of
the system is just determined by the largest, critical cluster.
As explained in chapter 4, the scaling properties of the percolation problem
are not so simple if one chooses instead to describe the system at the critical
threshold by a different measure, for instance the current in the random resistor
network. In this case, the current distribution is found to have multifractal scaling
properties, which in simple words means that for distinct regions in the distribution,
namely for each value of the current i, the increase of the quantity (n(i)} with L
is governed by a different exponent.
In order to check if also the distribution of local stress has a multifractal
nature, one can plot the rescaled distribution ln(n(f))/ In L as function of In |f |/ In L
for different system sizes L (fig. 7.9). For large L's the curves in this plot should
tend to the size independent f(a) spectrum, where, as discussed in chapter 4, the
scaling ansatz has been made that
(n(f)> ~ Lf{a) and |f| ~ L~a. (7.19)
Here the quantity f(a) represents the fractal dimension of the set of bonds expe-
riencing a force f and a is the singularity of this force, namely describes how the
value |f | goes to zero as the system size goes to infinity.
If simple scaling is followed, the function f(a) consists of a single point,
as for the case of the cluster size distribution in percolation, and the problem
is determined by one fractal dimension and one singularity. This is not the case
for the distribution of local forces just before rupture. Fig. 7.9 shows that indeed
the spectrum opens over a wide range of a's and therefore there is not a single
critical stress in the system. It is interesting to notice that, even for the case of
zero disorder in the thresholds, when the system is broken by a horizontal cut and
a single bond holds the system together, the spectrum f(a) is not reduced to a
single point. In particular, for both the fuse and central force model, the branch of
252 L. de Arcangelis / Randomness in breaking thresholds

f I 1 1 1 1—I i

r(I I I I
is I I
1 3 5 7 a

Figure 7.9 Rescaled log-histogram of the force distribution ln(n(f, L))/\nL versus In |f|/lnL in
the central force model just before the last bond breaks, for a unit external displacement and the
three different sizes L = 8,16, 24. This histogram should tend towards the multifractal spectrum
/(a) when the lattice size increases (from ref. 17).

the / ( a ) for small a values consists of a straight line delimited by the two points
(a = 1, / = 2) corresponding to the most probable value of the force scaling as
|f | ~ 1/L, and (a = 0, / = 0) for the last bond to be broken which experiences the
whole stress applied to the system. The sizes of the systems considered are still,
however, quite small and the statistical fluctuations too strong to bring further a
more quantitative analysis of the asymptotic f(a) spectrum within this approach.
Moreover this approach has log-corrections and it is therefore not very precise.
An alternative method to study the scaling properties of a given distribu-
tion is via its moments. One can define

M(q) = Σ lf l**(f) ~ L~Piq)- (7.20)


f

In particular, the zero-th moment, M(0), represents the mass of the current-
carrying part of the system. Just before failure the number of broken bonds is
rif ~ L 1 · 7 , therefore M(0) is expected to scale as L2.
The interest in defining the moments in eq. (7.20) is also in that the spec-
trum f(a) can be derived by Legendre transforming the exponents p{q), that is
by the scaling ansatz (7.19) and the relation dp(q)/dq = a (see chapter 4), one
obtains
f(a) = qa-p(q), (7.21)
which gives an alternative way to calculate the spectrum of fractal dimensions.
A preliminary check that can be performed to verify if the distribution has
multifractal scaling is to analyse the normalized moments mq = [M(q)/M(0)]1^q
as a function of L. If these quantities all scale with the same exponent no matter
the order q considered, then the moments in eq. (7.20) follow the constant gap-
exponent scaling. As mentioned before, this corresponds to a f(a) and a equal
4. Multifractality 253

10 30 100 10 30 100

Figure 7.10 Rescaled moments of the current distribution mq = [M(q)/M(0)]1'q for the fuse
model with the Weibull distribution with m — 2: (a) at the maximum of the characteristics and
(b) just before the system falls apart, in a log-log plot versus L. The symbols correspond to q = 1
( x ) , 2 ( Δ ) , 3 ( + ), 6 (a), 9 ( v ) and the zero moment M(0) ( O ) (from ref. 16).

to simple constants so that the spectrum of fractal dimensions consists of a single


point and the distribution is not multifractal.
Just before failure, when one bond holds the system together, the quantities
mq (fig. 7.10) do fan out, giving a clear indication for multifractality to hold for
the local stress distribution. However, if the same analysis is made not at the end
of the characteristics but at its point of maximum, the rag's do not fan out any
longer and regular scaling is followed. This result can be interpreted as follows:
in the first regime, where small isolated cracks are made, the system still behaves
as a macroscopically homogeneous object and only when correlations dominate
and a large spanning crack is formed, the system exhibits multifractal properties.
Physically, this means that the regions with very high local stress lie on a fractal
subset of the whole system and the fractal dimensions of these sets themselves
depend on how the stress scales with the system size (eq. (7.19)).
In order to further develop this analysis, for each value of q the asymp-
totic value of the critical exponent for the moment M(q) can be determined by
extrapolating the value of p(q) obtained from the finite size scaling plot to infinite
system sizes. In this way, one is able to eliminate the contribution to p(q) due to
the amplitude, which is determined to be different from unity, value found for the
positive moments in the percolation caseJ32^
In the fuse model, for instance, the exponents p(q) so determined just before
failure (fig. 7.11a) are found to be a monotonic function of the order q, which, for
q —> +oo, tends to a straight line of slope ~ 0.28 and vertical intercept at zero.
The behaviour shown is however very different from the linear dependence required
254 L. de Arcangelis / Randomness in breaking thresholds

r
k n—I—I—r—i—r f 1 1 I 1 1 1 ΓΊ
2.0
P(q)
0 1.6
/
1.2
/
//
\ :
" 1
f - 1
1
1 0.8 1
-8 -
-i
1

* 0.4
-12 I i 1 i 1 1 1 0 t\ . i i i 1 1

-<. 0 U 8 q D 1 2 3 (X

(a) (b)

Figure 7.11 Line of critical exponents p(q) versus q for the fuse model with a uniform distribution
(a). For the same model and distribution the spectrum of fractal dimensions /(a) (b) (from
ref. 19).

by the constant-gap scaling, since the curve sharply bends down for q values close
to zero. For the negative values of q it is rather difficult to accurately determine
the exponents which actually become quite large. Therefore it cannot be excluded
that the negative moments do not follow scaling but behave exponentially with L,
as proposed in the case of percolation^ (see also chapter 4).
As observed in fig. 7.10a, the moments M(q) at the maximum of the char-
acteristics follow instead a regular scaling. Therefore, the analogous determination
of the p(qYs in this case gives a simple linear dependence, p(q) = — 2 + g, where
the gap exponent is Δ = —1 for the scalar model.
Finally, the spectrum of fractal dimensions / ( a ) can be obtained once the
p(</)'s are known. The standard procedure consists in calculating the slope and
the vertical intercept for each point of the curve p(q). These two values will give
respectively the singularity a and the fractal dimension / ( a ) corresponding to a
given q value.
For the scalar model, the spectrum f(ct) calculated at the end of the char-
acteristics displays a typical bell-shaped behaviour (fig. 7.11b). The function has
a maximum, whose value, f(ct) = 2, corresponds to the "fractal dimension" of the
system. The location of such a maximum represents instead the singularity of the
most probable stress (current) in the lattice. Moreover, the left side of the spec-
trum, corresponding to the positive moments branch of the p(#)'s, goes to zero for
large g a t a value am{n ~ 0.28, which is the asymptotic slope of the curve p(q) as
q -> +oo.
These numerical results are consistent with the picture that at the final
point of breaking one fuse holds the system together, carrying the whole current
flowing in the network whose value dominates the positive moments M(q) for large
q. In such a configuration and for a unit external voltage applied, the current in this
most overstressed fuse is equal to the conductance G of the network. Therefore, the
singularity for the total current, a m i n , is expected to equal the critical exponent
4. Multifractality 255

of the conductance, p(2), since M(2) = G. The numerical results are indeed in
agreement with this prediction within error bars, since p(2) ~ 0.31.
The behaviour of the / ( a ) for higher a's is strongly affected by the sets
of bonds carrying a very small current or almost balanced, therefore the accurate
determination of the right side of the spectrum is at present beyond the available
numerical results.
In conclusion, the distribution of local stress at the very end of the rupture
process is wide and its scaling properties are not simply determined by a single
characteristic stress. More precisely, the system can be viewed as an ensemble of
sets of bonds, each characterized by a typical value of the local stress. Each set
then has its own scaling properties, related to the value of the singularity of the
corresponding characteristic stress.
These properties are characteristic of the point just before failure, whereas
during the whole disorder controlled regime the system follows constant gap scaling,
as if it were homogeneous. It is therefore the appearance of one macroscopic crack,
almost spanning throughout the system, that causes the richer scaling behaviour
of the system.
These multifractal scaling properties have been similarly found by anal-
ogous calculations in all three models for fracture both with wide and narrow
distributions of quenched disorder. Moreover, an analogous investigation on the
random fuse network with dilution disorder has also pointed out that the distribu-
tion of local current is very wide just before failure and follows multifractal scaling,
independent of the level of initial dilution consideredJ37l
At the final rupture point, the conductance (or elastic modulus) jumps
and the resistance (or compliance) jumps that occur when one bond is removed
from the network are also multifractal. This implies that the distribution of the
conductance and resistance jumps that occurred from the initial state to the final
rupture consists of two power-laws.^ This property should be easier to observe
experimentally than the multifractal character of distributions only at the final
rupture point.

Exponents p(q)
Q Fuse model Central force
0 -2.00 -2.0
1 -0.68 0.2
2 0.31 1.4
3 0.70 2.3
4 1.05 3.1
5 1.30 3.9
Table 7.2 Critical exponents for the moments M(q) of the current distribution (fuse model) and
the local force distribution (central force model) just before failure.

It has been proposed^ that the f(a) spectrum could be considered as a


256 L. de Arcangelis / Randomness in breaking thresholds

new tool to identify the universality class of a system, for which multifractality
holds. The spectrum would then play a very similar role to the one of critical ex-
ponents in ordinary critical phenomena. It would be therefore very nice to be able
to verify if the three different models for fracture proposed here do indeed generate
the same spectrum / ( a ) , or else the same set of exponents p(q). Unfortunately,
the numerical data so far available cannot give a definite answer to this question
(table 7.2). It would therefore be desirable to perform more accurate numerical
simulations on elastic models in order to verify if the universality found for the be-
haviour of the macroscopic breakdown quantities also holds for the local properties
of the system.
In the following chapter, we will see models for fracture with annealed
disorder. A connection between these models and the ones presented in this section
has been recently studied.^ In particular, it has been shown that it is possible to
reconstruct from an annealed fracture process, the corresponding distribution of
quenched disorder which generates the same patterns.

7.5 Discussion
Mechanical testings of materials probably started to be developed thou-
sands of years ago, when large scale wood or stone structures were first built. How-
ever, the first written record of a strength testing is due to Leonardo da V i n c i ^
(ca. 1500), who tested the tensile strength of iron wires of various lengths, by pour-
ing sand into a suspended basket. Since then the construction of more and more
sophisticated testing machines has made possible the analysis of the response to
an applied displacement of specimens of various materials, shapes and sizes.
In particular, although the first complete stress-strain curve of the kind
shown in fig. 7.2 was already known for concrete fifty years ago, some other prop-
erties needed the development of the so-called stiff, servo-controlled testing ma-
chines to be detected. This is the case for the classification of the fracture process
in different materials into two distinct categories. From the analysis of the com-
plete stress-strain curves for a variety of rocks, Wawersik^ in 1968 concluded that
rock failure is stable when energy must be added to the rock to promote further
failure, and unstable when energy must be instead extracted from the rock to avoid
violent collapse. He suggested furthermore that the stable or unstable failure under
uniaxial compression could be classified into two categories, class I and class II,
accordingly (fig. 7.12).
The so-called failure of class II, experimentally found in materials such as
basalt or granite, presents then the same behaviour obtained from numerical simu-
lations of systems with very small quenched disorder and sufficiently large system
size. As discussed in section 7.3, when the distribution of breaking thresholds is
very narrow, for instance the Weibull distribution with m = 5,10, the stress-strain
characteristics bend back toward smaller values of the strain after reaching the
catastrophic regime. The final failure is then attained by one macroscopic crack
5. Discussion 257

360
Stress
(MPa)
300

240

160

120

60

w
0 0.1 0.2 0.3 0.1, 0.5
Strain (%)

Figure 7.12 Complete compressive stress-strain curves for various rocks (after ref. 43). 1-3 =
class I: 1 = charcoal grey granite I; 2 = Indiana limestone; 3 = Tennessee marble I. 4-6 = class
II: 4 = charcoal grey granite II; 5 = basalt; 6 = Solnhofen limestone.

spanning the system, with no diffuse damage created.


This is the case experimentally detected for failure of class II. In many cases
the class II behaviour, in fact, is caused by non-uniform failure of the specimen,
non-homogeneously loaded: as one region of the sample experiences a strong stress
and fails, the rest of the specimen remains intact and is elastically loaded and
unloaded. This behaviour can actually be derived by means of a simple model,
which shows how the shape of the stress-strain characteristics changes when the
size of the specimen, L, is made increasingly larger than the size of the damaged
region, 5, kept fixed in the modelJ43^
Proceeding further from this first encouraging agreement with the exper-
imental data, one would like to verify next some of the universal scaling laws
numerically found for the breaking characteristics. Even if it is a well known fact
that the stress-strain curve crucially depends on the size and shape of the specimen
and extensive testings have been made in this direction, unfortunately the existing
data do not allow to make any statement on scaling.
To verify the indication for universality as found in the numerical simula-
tions is indeed a crucial issue. The concept of universality is extremely powerful in
critical phenomena: it allows to catalogue the physical behaviour of various systems
into few classes. Each class may gather very different physical systems, such as for
instance uniaxial ferromagnets and binary mixtures, and each class is characterized
by some exponents describing the critical behaviour. The holding of universal laws
in fracture would then allow to forget the microscopic details of the structure of
258 L. de Arcangelis / Randomness in breaking thresholds

different materials and to give a comprehensive, more general description of the


process.
The multifractal scaling properties are somehow more delicate to analyse.
Photoelastic experiments combined to image treatment techniques might help to
obtain a qualitative behaviour for the distribution of local stress in the solid. How-
ever, the evaluation of the spectrum of fractal dimensions requires high accuracy for
the local stress data and therefore needs more sophisticated techniques, as already
tested in different contexts such as, for instance, aggregation experimentsJ44^
On the other hand, better numerical simulations should also be performed
in order to confirm the universal properties of the spectrum f(ct) calculated for
different models. Especially for the central force and beam model more accurate
data are needed and the simulation should be pushed to larger system sizes. This
task probably requires the development of more efficient and performing numeri-
cal techniques, since the whole numerical work presented in the previous sections
already involved hundreds of hours of cpu on machines like the IBM 3090 or the
Cray XMP.
Finally, the three-dimensional case, more relevant from the experimental
point of view, should be more carefully analysed. Preliminary d a t a ^ on the central
force model seem to indicate that universality holds for the stress-strain character-
istics also in this case. However, it is of great interest to study the scaling properties
also for the scalar and beam model on three-dimensional lattices. Moreover, the
investigation of the distribution of local stress in the three-dimensional case may
lead to a deeper understanding of, for instance, the creation and evolution of faults
with planar structure, observed experimentally to be the major cause for failure in
real materials.

References
1. H. Liebowitz ed., Fracture, vols. I-VII, (Academic Press, New York, 1984)
2. R. Englman and Z. Jaeger eds.,Fragmentation and Flow in Fractured Media,
Ann. Israel Phys. Society 8, (Adam Hilger, Bristol, 1986)
3. E. Louis, F. Guinea and F. Flores , Fractals in Physics, L. Pietronero and E.
Tosatti eds. (Elsevier, Amsterdam, 1986), p.177
4. E.L. Hinrichsen, A. Hansen and S. Roux, Europhys. Lett. 8, 1 (1989)
5. P. Meakin, Thin Solid Films 151, 165 (1987)
6. Y. Termonia, P. Meakin and P. Smith, Macromolecules 18, 2246 (1985)
7. L. de Arcangelis, S. Redner and H.J. Herrmann, J. Physique Lett. 46, L585
(1985)
8. P.M. Duxbury, P.D. Beale and P.L. Leath, Phys. Rev. Lett. 57, 1052 (1986);
P.M. Duxbury, P.L. Leath and P.D. Beale, Phys. Rev. B 36, 367 (1987)
9. A. Gilabert, C. Vanneste, D. Sornette and E. Guyon, J. Physique 48, 763
(1987)
References 259

10. P.M. Duxbury and P.L. Leath, J. Phys. A 20, L411 (1987); P.D. Beale and
P.M. Duxbury, Phys. Rev. B 37, 2785 (1988)
11. P.D. Beale and D.J. Srolovitz, Phys. Rev. B 37, 5500 (1988)
12. H. Takayasu, Phys. Rev. Lett. 54, 1099 (1985)
13. M. Sahimi and J.D. Goddard, Phys. Rev. B 33, 7848 (1986)
14. B. Kahng, G.G. Batrouni, S. Redner, L. de Arcangelis and H.J. Herrmann,
Phys. Rev. B 37, 7625 (1988)
15. S. Roux, A. Hansen, H.J. Herrmann and E. Guyon, J. Stat. Phys. 52, 251
(1988)
16. L. de Arcangelis, A. Hansen, H.J. Herrmann and S. Roux, Phys. Rev. B 40,
877 (1989)
17. A. Hansen, S. Roux and H.J. Herrmann, J. Physique 50, 733 (1989)
18. H.J. Herrmann, A. Hansen and S. Roux, Phys. Rev. B 39, 637 (1989)
19. L. de Arcangelis and H.J. Herrmann, Phys. Rev. B 39, 2678 (1989)
20. F. Family, Y.C. Zhang and T. Vicsek, J. Phys. A 19, L733 (1986)
21. H.J. Herrmann, in Random Fluctuations and Pattern Growth: Experiments
and Models, H.E. Stanley and N. Ostrowsky eds., (Kluwer Academic Publish-
ers, Dordrecht, 1989)
22. L. Niemeyer, L. Pietronero and H.J. Wiesmann, Phys. Rev. Lett. 52, 1033
(1984)
23. J.P. Straley, Phys. Rev. B 15, 5733 (1977)
24. S. Feng and P.N. Sen, Phys. Rev. Lett. 52, 216 (1984)
25. S. Roux and E. Guyon, J. Physique Lett. 46, L999 (1985)
26. W. Nowacki, Theory of Micropolar Elasticity, (Springer-Verlag, Udine, 1972)
27. P.L. Leath and W. Tang, Phys. Rev. B 39, 6485 (1989)
28. S. Roux and A. Hansen, to appear in Europhys. Lett. (1989)
29. G.G. Batrouni and A. Hansen, J. Stat. Phys. 52, 747 (1988)
30. W.I. Newman and A.M. Gabrielov, preprint (1988)
31. S. Arbabi and M. Sahimi, preprint (1989)
32. L. de Arcangelis, S. Redner and A. Coniglio, Phys. Rev. B 31, 4725 (1985);
ibid. 34, 4656 (1986); L. de Arcangelis, A. Coniglio and S. Redner, Phys. Rev.
B 36, 5631 (1987)
33. R. Rammal, C. Tannous, P. Breton and A.M.S. Tremblay, Phys. Rev. Lett.
54, 1718 (1985)
34. G. Paladin and A. Vulpiani, Phys. Rep. 156, 147 (1987)
35. D. Stauffer, Phys. Rep. 54, 1 (1979)
36. B. Fourcade and A.M.S. Tremblay, Phys. Rev. A 36, 2352 (1987); G.G. Ba-
trouni, A. Hansen and S. Roux, Phys. Rev A 38, 3820 (1988)
37. Y. Li and P.M. Duxbury, Phys. Rev. B 40, 4889 (1989)
38. S. Roux and A. Hansen, in Disorder and Fracture, J.C.Charmet, E. Guyon and
S. Roux eds., (Plenum, New York, 1990) in press
39. L.P. Kadanoff, in On Growth and Form, H.E. Stanley and N. Ostrowsky eds.
(Martinus Nijhoff Publ., Dordrecht, 1986)
260 L. de Arcangelis / Randomness in breaking thresholds

40. A. Hansen, E.L.Hinrichsen, S. Roux, H.J. Herrmann and L. de Arcangelis,


preprint (1989)
41. L. Da Vinci, Testing the strength of iron wires of various lenghts, in Notebooks;
also mentioned in W.B. Parsons, Engineers and Engineering Materials in the
Renaissance, (Williams and Wilkins, Baltimore, 1939), p. 661
42. W.R. Wawersik, Detailed analysis of rock failure in laboratory compression
tests, unpublished thesis, University of Minnesota, Minneapolis, Minn. (1968)
43. J.A. Hudson, S.L. Crouch and C. Fairhurst, Eng. Geology 6, 155 (1972)
44. J. Nittmann, H.E. Stanley, E. Touboul and G. Daccord, Phys. Rev. Lett. 58,
619 (1987)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
©Elsevier Science Publishers B.V. (North-Holland), 1990 261

8. Dielectric breakdown and single crack models

Jänos Kertesz*

Pattern formation has become a rapidly developing field during the last
ten years or so; especially the phenomena of Laplacian growth have stimulated
much activity. The reason is at least twofold: rapidly increasing computer capaci-
ties including new dimensions of image processing provide formerly unimaginable
tools to approach these problems as well as recent developments in theoretical
physics like elaboration of concepts of fractal geometry and of microscopic solv-
ability have made it possible to handle some of the relevant nonlinearities. The aim
of the present article is to summarize some recent results relevant to single crack
propagation and to give a survey of the corresponding lattice models.
It seems advisable to learn about scalar Laplacian growth before going to
the vectorial problem of crack propagation which is at least from the numerical
point of view definitely more difficult. Therefore we start with an overview of the
physics of viscous fingering or dendritic crystal growth and with introducing the
basic computer model, diffusion limited aggregation (DLA). Concepts like instabil-
ities due to moving boundaries and fractal geometry will also be briefly discussed.
The branching structure of some crack patterns reminding to DLA-patterns
on the one hand and the analogies between the corresponding equations on the
other one lead physicists to transfer the methods and ideas used in Laplacian
growth to the field of fracture. Similar to the scalar case, the nonlinearity in slow
crack propagation stems from the moving boundary. Linear stability analysis is a
classic tool to obtain information about the early stage of the growth and about
the sensitivity of developing new modes with respect to perturbations. The first
results for crack propagation are already available.
Much of the work carried out by statistical physicists has been devoted to
computer simulations of several related discrete models which can be considered as
generalizations of DLA. Different physical situations and different approximations
resulted in a variety of models; those to be treated here have in common that they
all lead to single cracks. The main interest of physicists has been directed to the
t Institute for Theoretical Physics, University of Cologne, Zülpicher Str. 77, D-5000 Köln 41,
FRG; on leave from Institute for Technical Physics, H-1325 Budapest, Hungary.
262 J. Kertesz / Dielectric breakdown

//////////////////////

777777777777777777777Γ

Figure 8.1 Schematic picture of viscous fingering in a linear Hele-Shaw cell (see ref. 4). The
cell consists of two parallel horizontal glass plates separated by a distance b and filled with some
high viscosity fluid (often oil). At one of the open ends a fluid with low viscosity (e.g. air) is
pressed into the cell. After some transients a finger with a characteristitic λ = i/L ratio develops
where λ depends on the separation 6, the viscosity ratio and on the pressure. The z direction is
perpendicular to the plates.

fractal geometry of the patterns. Similar to the case of scalar growth problems the
crack patterns are also sensitive to noise and anisotropy. A way of controlling the
interplay between these two effects is noise reduction. This method allows one to
reach systematically the deterministic limit from the noisy case and to take into
account physically relevant memory effects.
A short discussion and a summary of open problems is given at the end of
this review.

8.1 Interfacial pattern formation


8.1.1 Laplacian growth
Viscous fingering in a Hele-Shaw cell^1-^ is a show case example of pattern
formation via interfacial growth. The Hele-Shaw cell consists of two glass plates of
width L separated by a distance b <C L and filled with oil (or some other viscous
liquid). The driving force is the applied pressure by which air (as a fluid with low
viscosity) is pressed into the cell from one of the edges pushing the oil out (fig. 8.1).

The starting point of the theoretical description is the linearized Navier-


Stokes equation of fluid motion:^

ηΔν = Vp, (8.1)

where η is the viscosity, v the velocity of the liquid and p the pressure. Due to the
symmetry of the geometrical setup, the solution is assumed to have the form
3 / 4.2 \
v(x, y, z) = - I 1 - -^ \ v„(:r, j/), (8.2)
1. Interfacial pattern formation 263

i.e. the problem can be considered as two-dimensional and it can be integrated out
in the z direction, perpendicular to the plates. Eq. (8.2) contains the boundary
conditions on the plates (v = 0) and the coefficient 3/2 is chosen so that the mean
over z of the velocity is equal to v\\(x,y).
Substituting eq. (8.2) into eq. (8.1) and taking into account that b <C L
one gets an equation which has the form of Darcy's law of fluid motion in porous
media: ^
v„ = - ^ V p . (8.3)
Assuming incompressibility, Vv = 0, we get the Laplace equation
Ap = 0 (8.4a)
in the domain V occupied by the fluid at time t. The boundary condition at the
liquid-gas interface r(t) is a balance equation:

ΡΓ = Pgas - 7«> (8-4b)

where pg3LS is constant if the viscosity of the gas is neglected. 7 is here the surface
tension and κ is the curvature of the interface. Since the curvature in the z direction
can be considered as independent of x and ?/, eq. (8.4b) can be integrated out in
the vertical direction. The normal velocity of the interface is vn = vyn, and for
the pressure at infinity we have p —> xv^ where v^ is the velocity of the liquid far
from the interface.
Introducing the dimensionless variable φ = (ρ9α8 — 7^± — p)b2 /{l^v^L)
with K_L being the mean perpendicular curvature, the full set of equations is:^1-4^
Δφ(χ, y) = 0, for (x, y) G V{t), (8.5a)
φΓ = α0κ, for (x,y) G Γ(ί), (8.5b)
vn = nV0, for (x, y) G Γ(ί), (8.5c)
where d0 = (b/L)2^f/(12ην^) is a constant proportional to the surface tension, (κ
in eq. (8.5b) is the curvature parallel to the plates.) Eqs. (8.5) represent a moving
boundary problem; eq. (8.5c) introduces the nonlinarity which is essential in order
to obtain pattern formation.
- Gradient governed Laplacian growth as formulated in eqs. (8.5) describes
a broad class of pattern forming phenomena, at least in some approximation. In
addition to viscous fingering, the following physical systems are worth mentioning
(in some cases the diffusion operator (dt — constΔ) replaces the Laplacian Δ in
eq. (8.5a) and the boundary conditions may also be more complicated):
- Dendritic crystal growth^~3^ from the melt where the driving force is
undercooling. The equation corresponding to eq. (8.5a) describes heat diffusion
and the boundary conditions are given by the Gibbs-Thomson relation and en-
ergy conservation. Dendritic growth from oversaturated solutions can be treated
in analogous terms.
264 J. Kertesz / Dielectric breakdown

- Electrodeposition^ of metal ions from a solution. The Laplace equation


is valid here for the electric potential and the driving force is the applied voltage.
The boundary conditions are determined in part again by particle conservation
and are more complex than in eq. (8.5).
Figure 8.2 contains some characteristic patterns grown in Laplacian pro-
cesses. We see that there are three types of patterns:^ dense branching (8.2 a, d,
g), open branching (8.2 b, e, h) and dendritic (8.2 c, f, i) structures. The former two
are characterized by splitting tips and the last one by stable propagating tips. It
depends on the physical conditions which of the possible patterns will emerge and
the morphological transitions between them constitute a current field of research.^

Why do eqs. (8.5) lead to pattern formation? As we shall see in the next
section, the originally symmetric solutions become unstable because of the non-
linearity introduced by the moving boundary condition (8.5c). The qualitative
explanation for the instability is that in gradient governed growth a protuberance
due to fluctuations will feel larger gradient resulting in faster growth i.e. increased
gradients and so on. This is the so-called Mullins-Sekerka^ or Saffman-Taylor
instability.^ (Clearly, changing the sign of the driving force removes the instabil-
ity.) The surface tension is a stabilizing factor; it introduces a lower cutoff and
perturbations of shorter wavelengths than this cutoff will die away.
8.1.2 Linear stability analysis and mode selection
In gradient governed Laplacian growth the moving boundary introduces a
nonlinearity which in turn may lead to instabilities and finally to pattern formation.
The linear stability analysis^ makes these considerations more quantitative. Let us
take a Hele-Shaw cell with the geometry of fig. 8.1 and represent the boundary by
a single valued function r(y, t) giving the position (x = r(y, t), y) of the interface.
The solution of eqs. (8.5) in the absence of perturbations is a straight front moving
with constant velocity r0(y,t) = xQ + 1 and a field φ0 = x — r0(y,t). Now a small
amplitude sinusoidal perturbation is put onto the interface

r(y,t) = r0(y,t) + 6k(t)eikv (8.6a)

and the field is expected to have the form φ = φ0 + φκ with

0fc = 6 fe (i)e^- fc « (8.6b)

Substituting this ansatz into (8.5b) and using κ « d2T/dy2 one gets

ek = 6k(l - d0k2). (8.7)

Then from (8.5c) the following equation for δ follows:

dtSk(t) = 1*1(1 - dok3)6h(t) (8.8)


1. Interfacial pattern formation 265

Figure 8.2 Pattern formation in crystallization (a, b, c), viscous fingering (d, e, f), and elec-
trodeposition (g, h, i). Three types of morphologies can be observed: dendritic (a, d, g), open
branching or fractal (b, e, h), and dense branching patterns (c, f, i). This set of pictures is repro-
duced from ref. 5. (a) AlGe crystal colony grown in an amorphous filmj9' (b) crystallization in an
amorphous GeSe2 thin film/10! (c) snow flakeJ11' (d) viscous fingering in a nematic liquid crystal-
air systemJ 12 ^ (e) water injected into a non-newtonian polymer fluidJ13^ (f) viscous fingering in
a radial Hele-Shaw cell (glycerine-air); in one of the plates a trianglar mesh was groovedj 14 ^ (g)
electrodeposition of Zn from an aqueous solution of ZnS(>4 in a quasi-twodimensional cell with
voltage - 6 V and concentration 0.01 Mj 1 5 l (h) 2 M J 1 6 ^ (i) 0.03 M ^
266 J. Kertesz / Dielectric breakdown

with the solution 6k(t) = exp[a;(fc)£]. ω > 0 means exponential growth, i.e. instabil-
ity. As anticipated in the previous point, the surface tension (do) has a stabilizing
effect: for wavenumbers k > l/y/do the modes decay exponentially (ω < 0).
A similar calculation can be carried out in the radial geometry (see e.g.
ref. 6). For the radial Hele-Shaw cell, where the gas is introduced through a central
hole into the system, this analysis leads to the following picture: As long as the
pressure is smaller than a threshold value given by the surface tension and the
radius of the starting circular pattern, the bubble shrinks. For larger pressures
the bubble grows and the interface will be unstable against perturbations. It is
convenient to choose the perturbations in this case as Amr~TneUrnt cos(ra<^) where
r is the distance from origin, φ the polar angle and Am an infinitesimal amplitude.
The larger the starting radius (or the applied pressure at a fixed radius) is, the
more m-modes are instable (have positive ωτη values). For large radii the relation
k ~ m/r connects the radial and linear geometries.
Spatial patterns develop when the symmetric solution of eq. (8.5) looses its
stability. The experiments show that there are several types of growing patterns
(fig. 8.2). An apparent difference between them is that while the dendritic shapes
(8.2 a, d, g) exhibit stable propagating tips, there is no stable microscopic shape
in the other cases. The tips of dendritic crystals grow with a constant velocity in a
well defined direction and have a radius of curvature which is uniquely determined
by the circumstances of the growth. Similarly, viscous fingers in a linear cell have
a well defined relative width λ. In the branching morphologies such modes cannot
be observed.
Historically, the first step towards the understanding of the mode selec-
tion was the treatment of the problems for the case of zero surface tension. For
the two-dimensional dendritic crystal growth the so called Ivantsov parabola,^
for the linear Hele-Shaw cell the Saffmann-Taylor fingers^ are the corresponding
solutions. These have in common that the equations determine a continuum of
curves in contrast to the observed unique selection of the modes and that they are
unstable against perturbations. This is not surprising because, due to the absence
of surface tension, a necessary characteristic length is missing and the system is
sensitive to any sort of perturbations.
In spite of some early expectations, the introduction of the surface tension
alone does not give yet the stable dendritic tip. This is nicely demonstrated on the
radial viscous fingering p a t t e r n ^ (fig. 8.3).
The pattern develops in the following way: first a symmetric, circular bubble
grows. When the radius becomes large enough (as can be calculated from linear
stability analysis), bumps start growing and become fingers. These increase and
thicken until the radius of curvature at their tips will again be too large - then the
tip of the finger splits. One of the new branches will eventually screen the other one.
Thus, the pattern is essentially determined by the sequence of these instabilities
due to tip splitting and screening.
The situation is different if in addition to the surface tension some kind
1. Interfacial pattern formation 267

Figure 8.3 Viscous fingering in a radial Hele-Shaw cell. The low viscosity fluid is introduced
through the central hole. (From ref. 18.)

of anisotropy is present in the system. In the linear Hele-Shaw cell, where the
geometry introduces the anisotropy, a stationary mode emerges after some transient
time (see ref. 3 and fig. 8.1). The relative width λ = i/L depends on the surface
tension and is constant in time for fixed driving force. Similarly, in crystal growth
the surface tension introduces anisotropy and the propagating tip of a dendrite has
a constant radius and velocity: The crystalline anisotropy stabilizes the tip.
The necessity of anisotropy in tip stabilization was shown mathematically
by transforming the problem of mode selection to a nonlinear eigenvalue equation
(we refer the interested reader for a detailed review and for more literature to
ref. 3). The strategy is the following: Assume the asymptotic shape of the pattern
far from the tip. Substitute this form into the equations of motion and build up the
solution numerically. It turns out that for fixed conditions there is a discrete set of
shape parameters and velocities (eigenvalues) which do not lead to a cusp at the
tip. Linear stability, now carried out around the propagating eigenfunction, shows
that the fastest one will be selected from these modes. The difficulties in solving
the problem of mode selection have their origin in the fact that both surface tension
and anisotropy turned out to be singular perturbations and both are needed for tip
stabilization. The mechanism by which the mode is selected is called microscopic
solvability.^
In the presence of tip splitting and screening the disordered open structure
of figures 8.2 b, e, h is expected. Under some circumstances a stable propagating
front of branches can be observed (fig. 8.2 a, d, g). By now it is not clear what are
the selection principles between these two morphologies, but randomness definitely
favours the former one.
8.1.3 Crack formation as a moving boundary problem
What happens if the fingering experiment is carried out in a viscoelastic
medium? For slow processes, when the characteristic time L/v is much larger than
268 J. Kertesz / Dielectric breakdown

Figure 8.4 Change from viscous fingering to cracking. Water was injected into a bentonite/water
paste at concentrations a) 0.08 and b) 0.20. (From ref. 20.)

the Maxwell relaxation time TM = ^/2// (μ is the shear modulus), the response is
viscous, i.e. the fluid flows. In the opposite limit, Ljv <C TM breaking phenomena
may happen. Thus fingering in a viscoelastic medium can provide a bridge between
the two kinds of pattern formation.
The crossover from fingering to crack propagation has been observed in
clay suspensions by Van Damme's group (see ref. 19 and section 2.6 of this book).
The non-Newtonian character of the flow and in some cases the miscibility of the
fluids used in the experiments make the observations difficult to interpret; however,
the qualitative trend is clear. There is a morphological transition of the patterns
(fig. 8.4) accompanied with a dramatic increase of the speed of propagation as
a function of increasing density (and therefore TM) of the substance. The fractal
dimension (to be discussed in subsection 8.2.1) changes from 1.7 to 1.4.
When increasing the density of the viscous fluid or the injection pressure
the branching angle changes from ~ 30 to ~ 90 degrees and the "fingers" become
cusp-like. These observations (changes in the speed of propagation, in the branching
angle, in the shape of the tips and in the fractal dimension) indicate that the
mechanism of pattern growth becomes different at the pressure ρτ ~ μ.
For p < p r eq. (8.5) is essentially valid (if non-Newtonian effects are dis-
regarded) and gradient governed growth with the characteristic tip splitting insta-
bility determines the morphology where an effective surface tension introduces the
lowere cutoff. In the opposite limit the equations of elasticity should describe the
system, and the equation for the moving boundary eq.(8.5c) should be replaced
by some threshold-type breaking criterion. Therefore the effective surface tension
does not play anymore the role of determining the characteristic curvatures and
the tips develop cusps. The tip splitting mechanism is replaced by side branching
1. Interfacial pattern formation 269

due to reaching the threshold stress somewhere behind the tip: This leads to the
change in the branching angle. 1
Usually energy arguments are used^ in order to decide whether a void in an
elastic medium becomes a propagating crack. However, the problem of crack prop-
agation itself constitutes a moving boundary problemJ23'24! Instead of the Laplace
equation we have now another elliptic differential equation, namely the Lame equa-
tion for the displacement vector u:

dadßUß + (1 - 2a)d2ßua = 0, (8.9a)

where σ is the Poisson ratio. Eq. (8.9a) expresses the vectorial character of the
process (cf. chapter 1). On the free boundary the condition

σ± = 0, (8.9b)

should be considered with σχ being the stress perpendicular to the surface of the
void. The condition for the moving boundary is different from eq. (8.5c), according
to the threshold-like behaviour:

vn oc σ\\ — ac (8.9c)

for σ\\ > ac and vn = 0 otherwise, where σ\\ is the stress parallel to the surface of
the crack and ac is the fracture stress threshold. The set of equations (8.9) is the
fracture counterpart of eqs. (8.5) of Laplacian growth. As there is no stabilizing
effect for σ\\ > ac it is not surprising that perturbations with all k wavenumbers
lead to instability as it can be demonstrated in a twodimensional calculation for a
circular voidJ24^ The observation that the tips of cracks have very large curvature
can be traced back to the very small scale perturbations which are always present
in materials subject to breaking.
Recently the question of steady state propagation of a crack in a viscoelastic
strip was addressed^ and shown to be simpler than the corresponding viscous
fingering problem. Starting from the viscoelastic strain-stress relation (which is
nonlocal in time, cf. eqs. (3.15) in chapter 3), assuming the geometry of fig. 8.5,
isotropy of the material, and a simple form for the cohesive stress near the tip as
a function of the crack opening displacement a unique propagating velocity and
a corresponding (Barenblatt-like^) tip-shape was obtained. The mode selection
problem for the crack propagation in a viscoelastic medium did not lead to the
subtleties discussed above for the Laplacian gradient governed growth.
1
It is worth mentioning that a similar sudden change in the mechanism of pattern growth
was observed in viscous fingering in a smectic liquid crystal^ 21 ! Around some threshold value of
the pressure the easy growth direction of the bubble in uniaxially preoriented smectics changed
by 90 degrees since the character of the displacements changed from anisotropic viscous flow to
solid-state-like defect mediated motion.
270 J. Kertesz / Dielectric breakdown

Mi | | 1
']_!_ ^ y

Figure 8.5 Schematic picture of a crack propagating at speed v in a viscoelastic strip of width
L under strain 6. (From ref. 23.)

8.2 Diffusion limited aggregation and dielectric breakdown


8.2.1 Fractal growth on computers
So far we have not considered noise explicitly. However, it is clear from
the nature of the instability that noise must play an important role in the growth
of patterns, especially in the case of small surface tension. The famous computer
model of diffusion-limited aggregation (DLA) by Witten and S a n d e r ^ is very
suitable to investigate this limit. The model is defined as follows: The cluster or
aggregate is initially a single particle. Another particle is launched at "infinity"
and undertakes a random walk until it hits the aggregate where it sticks to the
cluster. Then a new particle is started and the procedure is repeated, in principle
ad infinitum so that an aggregate forms similar to that shown in fig. 8.6. Here we
do not want to go into details of programming the modelJ1,4,27^
The connection to Laplacian growth becomes clear if we consider the two-
dimensional square lattice version of eq. (8.5a):

</>(*) = ϊ E < K x + a )· (8-10)


a
where the summation goes over the first lattice neighbours of x. Eq. (8.10) expresses
also the particle conservation if φ is interpreted as the probability of finding a
random walker at x.
The sticking of the particle corresponds to the moving boundary condition
in eq. (8.5c): vn oc ]£a</>(x + a) ~ V 0 r if ΦΓ — 0 (zero surface tension) and x is a
perimeter site of the cluster. Even in the absence of surface tension there is a lower
cutoff in the system introduced by the lattice constant. Of course, the rules can be
applied to other lattices or even without an underlying lattice - in the latter case
the lower cutoff can be identified with the size of the diffusing particles.
Niemeyer et al. defined a closely related computer model, the so-called
dielectric breakdown modelS28^ Here the Laplace-equation is solved numerically on
a lattice, the V^-'s are determined (i is the index of a perimeter site). Then one of
2. Diffusion limited aggregation and dielectric breakdown 271

1000 DIAMETERS

Figure 8.6 Off-lattice DLA cluster of size M = 50000. (Reproduced from ref. 27.)

the perimeters is occupied with probability

ft = |ν&|/Σΐν^·|, (8.ii)
j

according to eq. (8.5c). Now the Laplacian field has to be calculated again according
to the new boundary and the procedure is repeated many times until a large
aggregate is grown. The boundary condition expressed in eq. (8.11) is a special
case in the "77-model" where

A-=IW/EIW (8-12)
3

is taken. Even for η = 1 there is a slight difference between the boundary conditions
of DLA and of dielectric breakdown: while for the latter φ = 0 is taken on the
aggregate, φ = 0 already on the perimeter for DLA. However, this difference does
not seem to be relevant.
Figure 8.6 shows a typical off-lattice DLA cluster generated by MeakinJ27^
The above-mentioned sequence of tip-splitting and screening instabilities are clear-
ly determining the shape of the DLA clusters. It is also apparent that the noise,
which is an inherent part of the algorithms plays a crucial role in the growth.
However, the patterns show some regularity: they seem statistically self-similar
272 J. Kertesz / Dielectnc breakdown

and can therefore be described by Mandelbrot's fractal geometry (see refs. 1, 2,


29).
A fractal object has special scaling properties. Let us consider a fractal of
linear dimension L and cover it by boxes of size R. The number of boxes N needed
to cover the object is N ~ (R/L)D. Besides this so-called box counting method
there are other ways to measure D, the fractal dimension of the object. The number
of particles M can for instance be measured as a function of the distance R from
the origin leading to the following power law behaviour:
M ~ RD. (8.13)
If there is no ambiguity in using different lengths - as is the case for simple fractals
- R can be either the size of a box put on the fractal, the radius of gyration
Rg = [Ei>j( r i — Tj)2]1^2 o r a n v other reasonable length. The name dimension
becomes clear if we realize that for non-fractal, homogeneous objects D is just the
Euclidean dimension d. The origin of the non-integer D < d is that the pattern
contains holes of all sizes which lead to a size dependent density. Strictly speaking,
a fractal should be an infinitely large object and the above relationships are to be
understood as asymptotic laws. However, physically there are always finite lower
(particle size) and upper (system size) cutoffs and scaling like eq. (8.13) holds only
between these cutoffs. It is desirable, although not always possible, to satisfy that
the scaling region spans at least two decades.
Equation (8.13) is a result of the statistical self-similarity of the aggregates.
That means that the density-density correlation function c(r) is scale invariant:
c(br) = b~^c(r). This function equation has the solution c(r) = r'^ and by inte-
gration and comparison with eq. (8.13) we get ζ = D — d where d is the Euclidean
dimension of the space the aggregate is embedded in.
In some cases it is an affine transformation (and not a dilation) which
leaves the patterns (statistically) invariant, i.e. they are self-affineJ1'2'29! In order to
describe such objects one needs all the scaling exponents in the different directions.
Much computer time has been devoted to determine the fractal dimension
and to understand the structure of the DLA clusters. The most accurate value in
two dimensions is Ό — 1.715 ± 0.004 obtained^30' from the scaling of the radius
of gyration. This is close to the values which were measured on patterns grown in
systems represented by fig. 8.2b, e, h. In higher dimensions the formula^31! D =
(d2 + l)/(d + 1) seems to be a good approximationJ30l (Unphysical dimensions
d > 3 are of special theoretical interest.) The fractal dimension depends on η-S28^
the larger η is the more the growth on the tips is preferred leading to lower values
of D; for η = 0 we have D = d, while D —> 1 as η —> oo.
The single number D is not sufficient to characterize the full complexity of
DLA. As it is apparent from visual observation, the origin of the cluster plays a
special role in the cluster: around this point the dilation symmetry applies. This
observation is reflected in the tangential correlation function which measures the
decay of correlations at a fixed distance from the origin as a function of an angle.
2. Diffusion limited aggregation and dielectric breakdown 273

2.0 I I Γ

(a)

1.0

P
D-a
Ä-L. ■J 1 L
2 A 6 8 10
a
Figure 8.7 The multifractal exponents /(a) for two-dimensional DLA clusters. (Reproduced
from ref. 36.)

In contrast to isotropic fractals (like the infinite percolation cluster at threshold)


the tangential correlations in DLA decay faster than the radial onesJ32,33!
The part of the cluster where most particles arrive is called the active
zone}34} The width ξ of this active zone is defined as the mean deviation of the
growth probability distribution p(r) and £ ~ Mu' where M is the size of the cluster.
For a simple fractal one would expect i/ = \/D which expresses that all lengths
scale in the same way. At small to medium sizes this does not seem to be the case,
but the exponent v' turns out to depend slightly on M and probably approaches
1/D from below in the limit of M -> ooJ30^
The sites behind the active zone, deep in the fjords are rarely reached by the
randomly walking particles. The reason is well known from classical electrostatics:
It is the screening due to the elongated parts which makes the field small in the
fjords. It turns out that a finer subdivison of the cluster than active zone and dead
parts is of interest J35^ Let us consider the set of sites with growth probability p.
The strength of the screening is characterized by the exponent a defined
by
p(xR-a. (8.14)
The number of sites with the same p is denoted by n(p) and is expected to behave
like
n{p) oc β / ( α ) . (8.15)
The last equation expresses that the set of points with a given p (or a) is itself a
fractal. The function f(a) is nontrivial: there is an infinity of fractal dimensions
which are needed to characterize the system (fig. 8.7).
274 J. Kertesz / Dielectric breakdown

Figure 8.8 DLA-cluster grown with an anisotropic sticking rule on the square lattice. The
particles arriving at a growth site stick to the aggreagate with probability 1 if a left or right
occupied neighbour is present, otherwise the sticking probability is only 1/3. (Reproduced from
ref. 42.)

This phenomenon is called multifractality (see refs. 1, 2, 27 and also chap-


ters 4 and 7 of this volume). Of course, the whole multifractal formalism (Legendre
transforms, moments, Dqs) can also be applied to DLA: Σρ- ~ R^q~^Dq; Dq, a
and / are interrelated by (q — l)Dq = qaq - fq and aq = d/dq[(q — l)Dq] with
f(a) = fq(oLq). In addition to the generally valid relations (D = D0 = f0 = / m a x ,
D±oo = ct±oo, Di = <*i = / i ) we h a v e ^ 2DS = D and in two dimensions^ Dx = 1.
The extremely complex objects generated by the DLA algorithm are still
not fully understood and constitute an active field of research. Some recent works
are to be mentioned here. The first one concerns the screening deep in the fjords:^
it is possible that instead of the power law behaviour (8.14) a faster, exponential
decay sets in here. This would show up in a singularity in the f(a) curve. Another
work due to Vicsek et alJ40l is a recent careful numerical study of the geometrical
multifractality of DLA clusters. It seems that if the mass itself is used instead of the
growth probability as weight for the moments, the corresponding Dq's are again
depending on q. The most important consequence is, that fractal dimensions mea-
sured in different ways need not to be identical. Thus the "true" fractal dimension
(D0 — 1.73) was found to be slightly larger than that observed from the correlation
function or from the radius of gyration. Finally we mention the multiscaling picture
of the density profile^ g(x): Numerically g(x) a r~d+D(x)A(x) was found with a
varying fractal dimension D(x) where A(x) is a scaling function and x = r/Rg .

8.2.2 Anisotropy and noise


An interesting aspect of DLA is its reaction to external anisotropy which
can be either a direction dependent sticking rule or just the underlying lattice
structure. In the view of section 8.1 it is not surprising that DLA is sensitive to
anisotropy. In fact, an anisotropic sticking rule entirely changes the geometry of the
clusters: It leads to alongated cigar-shaped patterns which are not fractals because
the interior of the cluster becomes homogeneous^ (fig. 8.8).
The situation is more complex in the case of lattice anisotropy. Lattice
2. Diffusion limited aggregation and dielectric breakdown 275

Figure 8.9 The effect of noise reduction on DLA clusters with M — 400 and m — 2 (a), m = 20
(b) and m = 400 (c). (Reproduced from ref. 45.)

DLA clusters of small size are very similar to the off-lattice ones and the effect of
the underlying lattice becomes apparent only for larger sizes (e.g. for the square
lattice M « 105 particles).^
Why does the effect of the lattice show up so slowly? The reason is that the
fluctuations due to the random character of the walks act against the manifestation
of anisotropy. The actual shape of lattice-DLA clusters of a given size emerges from
an interplay between fluctuations and anisotropy.
The effect of this interplay can be investigated by the so-called noise re-
duction method which controls the fluctuations in DLAJ44,45] Instead of adding
the random walker to the cluster when it hits a growth site, one keeps counting
how many trajectories terminate at a given site. Only those sites are added to the
clusters at which the counter reaches a prescribed value m. A large m corresponds
to a low level of noise, m = 1 to ordinary DLA.
For a fixed number of particles an interesting morphological transition can
be observed as a function of m (fig. 8.9). Not too large noisy DLA clusters grown on
the square lattice are open branching structures with unstable tips like off-lattice
aggregates. If noise is reduced, the anisotropy due to the grid breaks through and
the clusters become cross-shaped with stable tips reminiscent of dendritic crystals.
For very large values of m even the sidebranches vanish and one recovers needles
growing out of a center. The transition from tip-splitting to stable tip can also be
seen in a multifractal analysis:^ the localized propagating stable tips appear as
single peaks in the n vs. lnp spectrum and the f(a) corresponding to the dimension
of the set with maximal growth probability becomes zero.
What happens if the size M of the aggregate is increased for a fixed value of
m? There is theoretical and numerical evidence^ that the asymptotic shape will be
the (fractal) dendrite (and not the needle). The critical size at which side branching
sets in on a square lattice goes as [log(ra)]3. Ordinary lattice DLA clusters (m = 1)
276 J. Kertesz / Dielectric breakdown

are expected to behave asymptotically like noise reduced ones: the lattice affects
the fractal dimension as seen in fig. 8.10 (D « 1.53 for the square l a t t i c e ^ ) . This
means that both anisotropy and noise are relevant parameters to the problem of
the asymptotic shape of DLA clusters.
The physical interpretation of noise reduction can be the following: Let us
imagine a square lattice where the bonds are made of pipes filled with a viscous
liquidJ48^ At the central node air is introduced, i.e. we consider a fingering experi-
ment in this system. The exact theoretical description would be the implementation
of eq. (8.5) to this geometry. However, a possible approximation is that the Laplace
equation is solved on the nodes of the square lattice and we keep record of how far
the air could advance in the pipes. If the Laplace equation is solved by the method
of random walks and the motion in the pipes is discretized into m steps we arrive
at the noise reduced DLA model.
Some features become clear from this physical implementation of the model.
Noise reduction introduces a sort of memory as compared to the m = 1 case: The
"solution of the Laplace equation" is carried out irrespective of the status of the
counters, but a node will have higher probability to be occupied if it has been
visited already many times. Furthermore, the limit m —> oo becomes a well-defined
deterministic problem.
The type of fluctuations we considered are generated during the growth
and are an example of so called annealed randomness introduced in chapter 5.
An interesting question is whether the different type of disorder, namely quenched
randomness leads to new effects, as it has been observed for critical phenomena.
Quenched randomness means here that the growth probability depends on random
variables which are fixed before the growth starts. Chen and Wilkinson^ supposed
in the above discussed model of pipes a random distribution in the radii of the
pipes. Studying the system both experimentally and by simulations they arrived
at the conclusion that the observed patterns are DLA-like with a fractal dimension
close to that of the DLA. Family et alJ49^ introduced a combination of quenched
and annealed disorder: the Laplacian field was considered on a lattice consisting
of resistors of random conductivity σ. The quantity p = σ(\/φ)η was calculated
where σ was randomly distributed in the interval [0,1]. The sites with maximal
p were added to the cluster. Fractal clusters were produced with D varying from
1.37 (η = 1/6) to 1.16 (η = 1/3).
Some important experimental work on quenched disorder should also be
mentioned here. A first example is viscous fingering in a network of pipes with a
random distribution of radii J50^ Chen and Wilkinson^ observed similar transitions
as shown in fig. 8.9: as the distribution of radii got sharper the patterns crossed
over from random fractals to dendrites and finally needles. This demonstrates that
annealed and quenched disorder have, at least qualitatively, similar effects. In ac-
cordance with this, Mäl0y et alJ51^ carried out viscous fingering in a Hele-Shaw
cell filled with small glass spheres and observed nice DLA-type patterns. Honjo et
alJ52l demonstrated that the tip-stabilizing effect of anisotropy can be suppressed
2. Diffusion limited aggregation and dielectric breakdown 277

CO

2
3

O
o
o
m

·· h CO

z
i
uz '■
3
*-> O
1 .

*> o
o
tf «£ <>
o
CD

< -·
-9»
M

V <

Figure 8.10 A coarse grained DLA cluster of 4 x 106 particles grown on the square lattice (a).
The last 5% of the particles is also displayed (b). (Reproduced from ref. 27.)
278 3. Kertesz / Dielectric breakdown

l/r1

y=i

i/=0

Figure 8.11 Schematic picture of the dielectric breakdown model (a) and its dual (b). (Repro-
duced from ref. 53.)

by quenched randomness: crystals grown on a rough plate showed tip splitting.

8.3 Single crack models on lattices


8.3.1 From dielectric breakdown to fracture growth
In this part we turn to two-dimensional lattice models of cracking. Some
important aspects have already been emphasized in former chapters of this volume.
Let us start with the dielectric breakdown model (cf. eq. (8.10)) on the square
lattice in the strip geometry, i.e. this time the cluster does not grow from a single
seed but between two parallel electrode plates at the top and the bottom on which
a constant voltage difference is imposed (fig. 8.11a). The lattice version of the
Laplace equation is solved in this geometry and, with probability proportional to
the local voltage drop, sites are joined to the aggregates that are attached to one
of the boundaries. In order to maintain the voltage of the clusters uniform, one
assumes that the sites in the cluster are superconducting in contrast to the rest of
the bonds which have some equal finite resistance.
The dual to this model (see fig. 8.11bJ53^ and also chapter 4 of this volume)
is defined in the following way: Current is driven through the system but this
time perpendicularly to the former direction of the average voltage drop. Now the
clusters consist of broken (infinite resistance) bonds and the probability of breaking
a bond adjacent to the clusters is proportional to the local current which is obtained
by solving the Kirchhoff equations (i.e. since all conducting bonds are equal, by
solving the lattice Laplace equation). The dual model defined in this way is in two
dimensions equivalent to the original one. (In higher dimensions the equivalence
does not exist anymore.)
When the cluster of broken bonds percolates, the system breaks into two
pieces. Therefore the dual dielectric breakdown (or the fuse model) is already a
toy fracture model;^53,54^ one should imagine that the bonds are broken because of
3. Single crack models on lattices 279

Figure 8.12 Three different possible definitions of neighbourhood on the triangular lattice.
Broken bonds are marked by dashed lines. (Reproduced from ref. 55.)

a horizontally applied stress. As in two dimensions, the dual dielectric breakdown


model is equivalent to the original one all the formerly obtained results can just
be taken over.
It is indicated in fig. 8.11 that the dielectric breakdown algorithm - and
consequently the fuse model - in the strip geometry results in a number of distinct
clusters one of which will finally span the system. In order to arrive at single
cluster models special restrictions are to be made, namely a single seed particle
should be placed on the lattice and during the evolution of the pattern growth
sites are only allowed in the neighbourhood of the cluster. Although this restriction
seems somewhat artificial, it can be motivated in some cases since the crack often
weakens the material. The definition of neighbours introduces some arbitrariness
and several different choices have been tried (strict selfduality on the square lattice
defines neighbourhood unambigously). Fig. 8.12 shows three possible definitions
on the triangular lattice as used in ref. 55.
8.3.2 Growth of cracks
In order to arrive at a real model for crack propagation one important
step remains: The elastic character of the medium should be taken into account.
Instead of the scalar Laplace equation its "vectorial counterpart", eq. (8.9a), the
Lame equation of elasticity^ for the displacement vector u is to be solved.
The simplest way of obtaining a discretized form of eq. (8.9a) is to suppose
that the bonds of the lattice consist of springs which can freely rotate around the
nodes. This is the so called central-force model^ (see chapter 4) where forces o
280 J. Kertesz / Dielectric breakdown

axis between the endpoints. This assumption is somewhat artificial: For example
the shear mode does not exist for the square lattice, therefore this model is usually
studied on the triangular mesh.
The full Born mode/'58' is a generalization which incorporates an additional
scalar term:

v = !(*i - **) Σ (to - u i) · fy)2 +1*» Σ Κ - "i)\ (8-16)

where V is the total elastic energy, f^ is a unit vector from site i to j , k\ and k2
are force constants. k2 = 0 gives the central force model and fc2 = &i reduces to the
scalar case of the fuse model J59' Although the Born model is unphysical since it is
not rotationally invariant, the investigation of the fractal behaviour of the cracks
in this framework can be of theoretical interest.' 60 '
Another, more realistic possibility of taking into account forces beyond the
central ones is to consider the bonds of the lattice to be elastic beams. This so
called beam mode/'61' has already been introduced in chapter 3 of this volume. The
basic point is that the energy of the bond is increased not only by elongating but
also by bending or shearing the beams, therefore local twist at the nodes according
to Cosserat elasticity (see ref. 62) is possible.
Once the elastic model is defined the cracking criterion has to be specified
(cf. eq. (8.11)). The criterion for local growth is theoretically not so clear as in
the case of Laplacian growth. Louis and Guinea'68' assumed the proportionality of
the breaking probability pi with the local force in analogy to DLA. For the central
force model this means
Α = Μ/ΣΙ*;Ι> (8·17)
3

where £,· is the local strain of the z-th spring. Just as for dielectric breakdown in
eq. (8.12), the "77-parameter" can be introduced' 55 ' with pi oc \δ{\η. This breaking
criterion has also been used for the full Born model.'60' The situation is somewhat
more complicated for the beams since here the breaking must also depend on the
local moments πΐ\ and ra2 occuring at the ends of the beam. The form'64'

Pi oc (f2 + qmax(\m1\, \πι2\))η (8.18)

is suggested by analogy to the von Mises yielding criterion for 77 = 1, where / is


the traction (compression) force and q parametrizes the affinity of the breaking
process to the bending mode. "Thermal activation" (see chapter 9) has also been
used for single cracks as a breaking criterion.'65'
The models are implemented on a computer in the following way. A regular
two-dimensional lattice (of size « 100 x 100) is defined. The proper external stress
or strain on the boundary is applied (shear, uniaxial stress, uniform dilation or
compression); periodic boundary conditions can be used in the allowed directions.
The bond in the middle of the lattice is broken and the system is relaxed to its
3. Single crack models on lattices 281

equilibrium (the breaking criteria and procedures model slow crack propagation.)
It seems that the most powerful relaxation procedure is the conjugate gradient^
method but simple overrelaxation has also been used. The next step after relaxation
is to choose one bond from the surface of the crack (where the neighbourhood has
to be defined before, see fig. 8.12). This bond is broken i.e. it is added to the
crack and the system is then relaxed with this new configuration. The procedure
is repeated until the computer facilities are exhausted. The grown cracks consist
of typically « 102 — 103 broken bonds. Averages over several ( « 10) runs are taken
in order to reduce statistical errors.

8.3.3 Fractal crack patterns


It is clear from the previous considerations that there is a variety of possible
cases depending on the model (fuse, central force, beam), on the applied stress or
strain (uniaxial stress, homogeneous dilation or compression, shear) on the neigh-
bourhood definition (cf. fig. 8.12) and on the breaking criterion. In the following
we shall try to summarize the most important findings in these families of models.
The most important general message is that patterns generated by the
above specified crack models are fractalsJ 55,60,63_65,67_69 l This is usually demon-
strated by measuring the radius of gyration of the crack pattern as a function of
the number of bonds forming the crack. The slope of the more or less linear log-log
plot is the inverse fractal dimension. The simulations are confined to rather small
systems, thus it is more appropriate to speak about effective fractal dimension. Es-
pecially in the light of the rather slow convergences observed for DLA, one expects
slight changes in the values of the effective fractal dimension when going to larger
samples.
Some of the results can be understood by analogy with the dielectric break-
down model. The fractal dimension D is sensitive to the parameter 77t55'68] and the
trend is that for probabilistic models D decreases with increasing 77. Increasing the
number of bonds in the definition of the neighbourhood seems^55! to increase D.
However, due to the limited size of the cracks it is hard to judge whether these
differences would survive for asymptotically large cracks.
The most extensively studied case is the central force model on the trian-
gular latticeJ55,63,66,68^ Figure 8.13 shows the patterns observed imposing shear and
uniaxial tension for η = lJ68J The simple calculation of D suggests different values
for the two cases (D « 1.42 for shear, D « 1.22 for uniaxial tension). However,
it is likely that the shear pattern fig. 8.13a should be considered as composed by
two arms each being similar to fig. 8.13b^67^ corresponding to the fact that a simple
stress can be decomposed as a sum of a traction and a perpendicular compression.
This argument was supported by the scaling analysis of the inertial tensor of the
cracking patterns where the two principial values were found to scale differently.
The larger exponent ( ^ 1.28)f67' is close and probably asymptotically equal to the
fractal dimension of the crack under uniaxial tension. The fact that two different
exponents are needed to describe the scaling behaviour^ shows that the crack
282 J. Kertesz / Dielectric breakdown

S = 750

Figure 8.13 Crack patterns generated in the central force model on the triangular lattice, η = 1,
neighbourhood I according to fig. 8.12, (a) shear, (b) uniaxial tension; 5 is the number of broken
bonds. (Reproduced from ref. 68.)
3. Single crack models on lattices 283

pattern is not simply self-similar, but rather self-affine. The isotropic dilation leads
to D « 1.35 (cf. table 8 . 1 ) . ^

77-parameter stress field neighbours D


dilation I 1.35
dilation II 1.51
dilation III 1.66
shear I 1.42
shear II 1.62
shear III 1.65
2 dilation I 1.12
2 dilation II 1.16
2 dilation III 1.45
2 shear I 1.17
2 shear II 1.49
2 shear III 1.40
Table 8.1 Effective fractal dimensions obtained from the central force triangular lattice model J55^
Strong deviations due to systematic and statistical errors are possible. The notation of the neigh-
bours refer to fig. 8.12.

A possible interpretation of the breaking criteria is stress corrosion, thus


memory effects according to preweakening of bonds due to stress are of physical
relevance. This provides further motivation^ for application of the method of
noise reduction as discussed in part 3. The method can be applied to the fracture
growth models in a straightforward wayJ70^ Counters are put now on the bonds
on the crack surface. After calculating the 6's, trials are made according to (8.17)
or (8.18) until one of the bonds is broken because it has been chosen m times.
Then a new equilibrium is determined but the values at the "counters" are kept.
The transition from random cluster to regular dendritic pattern was observed in
the central force model on the triangular lattice together with a decrease of the
fractal dimension, in analogy with the case of scalar DLA. However, no transition
to needles have been observed for m < 1000J70^
The infinite noise reduction limit was one of the deterministic rules applied
to the beam modelJ64^ Three possibilities were considered:
I. One just breaks the beam with the largest value of p from eq. (8.18).
II. One breaks the beam for which q0 = p + foP-i is largest where p_i is the value
of p that this beam had before the previous beam was broken; / 0 is a memory
factor.
III. On every beam of the lattice a counter c is put which is set zero at the very
beginning. Each time one has obtained the p's one calculates a = (1 — c)/p and
breaks the beam which has the smallest a namely a m i n · After the beam has been
broken, every counter is set to c = am\np-\- /ic_i where c_i is the value the counter
284 J. Kertesz / Dielectric breakdown

Figure 8.14 Deterministic crack pattern under external shear generated with short term memory
in the beam modelJ64^ The parameters are: Criterion II, q — 0, η = 0.7, /o = 1 and the size of
the square lattice L — 118. D = 1.25 for this pattern.

had before the breaking and / i is another memory factor. Criterion III corresponds
for / i = 1 to the limit of infinite noise reduction.
Physically the three breaking criteria introduced above correspond to three
different situations. Criterion I describes ideally brittle and the fastest rupture. Cri-
terion II contains a short time memory one would expect in cracks that propagate
with a finite velocity and produce strong local deformations at the tip of the crack
with very fast stress corrosion effects. Criterion III can be applied to situations of
relatively slow stress corrosion. The memory factors / 0 and / i measure the strength
of these time correlations and in criterion III the limit / i —> 0 gives criterion II
with /o = 1 whereas in criterion II the limit / 0 —» 0 gives criterion I.
Applying rule I to the shear case leads to a crack pattern which is not
fractal, but - in contrast to the analogous scalar case - it has branches. (A "DLA-
cluster" with an analogue rule would consist of two perpendicular straight lines.)
The reason is that cleavage tends to have the crack grow in the diagonal direction
while the bending mode favours a horizontal rupture.
In fig. 8.14 we see a crack grown using criterion II with q = 0, η = 0.7 and
a size L — 118. The fractal dimension D is determined by counting the number of
broken beams inside a box of length / around the first broken beam and plotting it
as a function of I in a log-log plot ("sand-box method"). D « 1.25 was obtained.
Changing the elastic constants just changes the opening angle of the crack
while an increase in η decreases the density of side branches and also the fractal
dimension (which is an effect due to the deterministic character of the growth).
The introduction of randomness results in an increase of the fractal dimension,^
in accord with the general experience.
3. Single crack models on lattices 285

(a)

~-<^^5^£L/S^

Figure 8.15 Cracks grown on a 60 x 60 square lattice under external shear and breaking beams
only by traction (q = 0).[64] (a) Criterion III, η = 1, /χ = 1. (b) Criterion II, η = 0.2, / 0 = 1. (c)
For comparison the morphology of cracking in aged (100 h, 750 K) Tin.5Mo6Zr4.5Sn78, tested in
0.6 M LiCl in methanol at 500 mV under increasing stress intensity.^73!

The difference to DLA is apparent from criterion I: a similar scalar growth


rule leads obviously to nonfractal DLA-clusers. It should be noted, that determin-
istic scalar Laplacian fractals have also been generated by using algorithms that
occupy growth sites where the gradient was larger than a threshold^71,72! (and not
just the maximal one as in our case). The origin of fractality on figures 8.14 and
8.15 is the competition between a global stress perpendicular to the diagonal and
a local stress that tends to continue a given straight crack due to tip instability.
Again we see the important role of the interplay of different directions which is
only possible in a truly vectorial model.
Figure 8.15a shows a crack grown using criterion III for q — 0, / = 1,
7? = 1 and L = 60. The physical situation is similar to that seen in criterion II,
only the fractal dimensions are higher. This case can be directly compared to DLA
where in the limit of infinite noise reduction needles, not fractals are predicted.
An experimental example of stress corrosion cracking is also displayed, to show
apparent similarities to the simulations. Although no quantitative comparison can
be made, it seems that at least some of the geometrically relevant features are
captured by the models.
286 J. Kertesz / Dielectric breakdown

8.4 Discussion and outlook

The main point of this chapter has been the introduction of the concept
of fractal geometry to the field of crack growth. In 1984 Mandelbrot et alJ74^ pub-
lished a paper with the title: "The fractal character of fracture surfaces of metals"
and since then some further experimental evidence has accumulated ranging from
micro- through macro- to mega-scale, from microcracks to the determination of
the fractal dimension of the San Andreas fault J75'76J
The discussed lattice models show that self-similar patterns with nontriv-
ial fractal dimension are produced by more or less realistic breaking procedures.
Important concepts could be taken over from the much better understood physics
of diffusion-limited or Laplacian scalar growth.
Probabilistic single crack growth processes lead to fractals in all consid-
ered cases. This is in accord with the corresponding experience for DLA. However,
the situation is much more complex for cracks: different boundary conditions lead
sometimes to different dimensions^55,68! (although this may finally turn out to be
a finite size effect). Simple short time memory as introduced for the determinis-
tic beam m o d e l ^ resulted in fractal clusters in contrast to the analogous scalar
models.
Several questions are going to be addressed in the near future. The growing
crack is expected to show multifractal behaviour with respect to the cracking prob-
ability; however, its conclusive analysis is not possible with the available computing
tools. Going to three dimensions is in principle straightforward but the computable
crack sizes would be even farther from "asymptopia" than in the two-dimensional
case. Due to the relaxation mechanism the growth of the cracks proceeds with a
speed much smaller than the sound velocity. For many processes this is an ac-
ceptable assumption but there are cases where it should be dropped and the time
dependent elastic equations should be taken into account. Another point is that the
lattice structure plays an obviously important role in the simulations. It may be
considered to mimic some intrinsic anisotropy; however, simulations on amorphous
graphs would also be interesting since in many cases fracture develops intergranu-
larly i.e. without underlying long range anisotropy.
It is obvious that in most physical situations the randomness has rather
quenched than annealed character. Quenched randomness means that the breaking
probability depends on the strength of the bonds which is a random variable fixed
prior to the beginning of the fracture. The investigation of the role of the quenched
randomness is expected to be continued since virtually all types of models can be
generalized to this physically important stuation.
The most important problem, on which research should concentrate is the
relation of the applied models to physical reality, the clarification of possible rel-
evance of fractal properties to materials science. So far it seems that the fractal
dimension serves as a classification criterion but we have not understood much
about the mechanisms leading to different D values and how the fractal dimension
References 287

of the fault is related to mechanical properties of the materialsS77^ Some first steps
towards understanding the physics of the models are the above mentioned picture
about the possible superposition of shear cracks from patterns produced under
uniaxial s h e a r , ^ the possible universal behaviour in the presence of noncentral
forces in the Born m o d e l ^ or the role of memory in producing deterministic frac-
tal cracksJ64^ However, the available model parameters allow practically for any
fractal dimension between 1 and 2, therefore there is urgent need in model calcula-
tions where quantitative relations to physical parameters can be established. And
last but not least: evaluation of experimental data using fractal concepts should
be continuedJ 74-77 ]
The serious problem with the lattice models discussed above is that simula-
tions are extremely time consuming. It took e.g. 4 hours CPU time on a
CRAY XMP to obtain the deterministic crack in fig. 8.14. The properties we are
looking for are supposed to show up asymptotically, i.e. for large sizes. The crack
patterns containing at most a few thousand broken bonds are to be compared to
DLA clusters of « 106 sites. Averaging over many samples should be carried out
for models with randomness. Clearly, we are far from being able to give definite
answers to the questions arisen. It is a real challenge of computer physics to find
much more efficient algorithms.

I would like to thank Lucilla de Arcangelis, Hans Herrmann and Stephane


Roux for their advice; thanks are also due to Paul Meakin and Tamäs Vicsek for
preprints. The support of the Alexander von Humboldt Foundation is gratefully
acknowledged.

References
1. T. Vicsek, Fractal Growth Phenomena (World Scientific, Singapore, 1989)
2. J. Feder, Fractals (Plenum, New York, 1988)
3. D.A. Kessler, J. Koplik and H. Levine, Adv. Phys. 37, 255 (1988)
4. D. Bensimon, L.P. Kadanoff, S. Liang, B. Shraiman and L. Tang, Rev. Mod.
Phys. 58, 977 (1986)
5. T. Vicsek and J. Kertesz, Europhys. News 19, 24 (1988)
6. J. Kertesz, in Random Fluctuation and Pattern Growth: Experiments and
models, ed. H.E. Stanley and N. Ostrowsky (Kluwer, Dordrecht, 1988) p. 42
7. W.W. Mullins and R.F. Sekerka, J. Appl. Phys. 34, 323 (1963)
8. P.G. Saffmann and G.I. Taylor, Proc. R. Soc. London A. 255, 312 (1957)
9. E. Ben-Jacob, Y. Godbey, N. Goldenfeld, J. Koplik, H. Levine, T. Mueller and
L.M. Sander, Phys. Rev. Lett. 55, 1315 (1985)
10. G. Radnoczy, T. Vicsek, L.M. Sander, D. Grier, Phys. Rev. A 35, 4012 (1987)
11. W.A. Bent ley and W.J. Humphreys, Snow Crystals (Dover, New York, 1962)
12. A. Buka, J. Kertesz and T. Vicsek, Nature 343, 424 (1986)
13. G. Daccord, J. Nittmann and H.E. Stanley, Phys. Rev. Lett. 56, 336 (1986)
288 J. Kertesz / Dielectric breakdown

14. E. Ben-Jacob, G. Deutscher, P.Garik, N. Goldenfeld, Y. Lereah, Phys. Rev.


Lett. 57, 1903 (1986)
15. Y. Sawada, A. Dougherty and J.P. Gollub, Phys. Rev. Lett. 56, 1260 (1986)
16. M. Matsushita, M. Sano, Y. Hayakawa, H. Honjo and Y. Sawada, Phys. Rev.
Lett. 53, 286 (1986)
17. G.P. Ivantsov, Dokl. Akad. Nauk SSSR. 558, 567 (1947)
18. L. Paterson, J. Fluid. Mech. 113, 513 (1981)
19. H. Van Damme, in The Fractal Approach to Heterogenous Chemistry, D. Avnir
ed. (Wiley, 1989) p. 199
20. E. Lemaire and H. Van Damme, C. R. Acad. Sei. Ser. II. 39, 859 (1989)
21. V.K. Horväth, J. Kertesz and T. Vicsek, Europhys. Lett. 4, 1133 (1987)
22. A.A. Griffith, Philos. Trans. Roy. Soc. A 221, 163 (1920)
23. M. Barber, J. Donley and J.S. Langer, Phys. Rev. A 40, 366 (1989)
24. H.J. Herrmann and J. Kertesz (unpulished)
25. G.I. Barenblatt, Adv. Appl. Mech. 7, 56 (1962)
26. T.A. Witten and L.M. Sander, Phys. Rev. Lett. 47, 1400 (1981)
27. P. Meakin, in Phase Transitions and Critical Phenomena, C. Domb and J.L.
Lebowitz eds., (Academic, New York, 1988) Vol. 12, p. 335
28. L. Niemeyer, L. Pietronero and H.J. Wiesmann, Phys. Rev. Lett. 52, 1033
(1984)
29. B.B. Mandelbrot, The Fractal Geometry of Nature (Freeman, New York, 1982)
30. S. Tolman and P. Meakin, Phys. Rev. A 40, 428 (1989)
31. M. Muthukumar, Phys. Rev. Lett. 50, 839 (1983)
32. P. Meakin and T. Vicsek, Phys. Rev. A 32, 685 (1985)
M. Kolb, J. Physique Lett. 46, L631 (1985)
33. B. Mandelbrot and T. Vicsek, J. Phys. A 22, L377 (1989)
34. M. Plischke and Z. Räcz, Phys. Rev. Lett. 95, 415 (1984)
35. T.C. Halsey, M.H. Jensen, L.P. Kadanoff, I. Procaccia and B. Shraiman, Phys.
Rev. A 33, 1141 (1986)
36. Y. Hayakawa, S. Sato and M. Matsushita, Phys. Rev. A 36, 1963 (1987)
37. T. Halsey, Phys. Rev. Lett. 59, 2067 (1987)
38. N.G. Makarov, Proc. London Math. 51, 1119 (1985)
39. J. Lee, P. Alstr0m and H.E. Stanley, Phys. Rev. A 59, 6546 (1989)
40. T. Vicsek, F. Family and P. Meakin (preprint)
41. C. Amitrano, A. Coniglio and M. Zannetti (preprint)
42. R.C. Ball, R.M. Brady, G. Rossi and B.R. Thompson, Phys. Rev. Lett. 55,
1406 (1985)
43. P. Meakin, R.C. Ball, P. Ramanlal and. L. Sander, Phys. Rev. A 35, 5233
(1987)
44. C. Tang, Phys. Rev. A 31, 1977 (1985)
J. Szep, J. Cserti and J. Kertesz, J. Phys. A 18, L413 (1985)
J.Nittmann and H.E. Stanley, Nature 321, 663 (1986)
45. J. Kertesz and T. Vicsek, J. Phys. A 19, L257 (1986)
References 289

46. C. Amitrano, L. de Arcangelis, A. Coniglio and J. Kertesz, J. Phys. A 21, L15


(1987)
47. J.P. Eckmann, P. Meakin, I. Procaccia and R. Zeitak, Phys. Rev. A 39, 3185
(1989)
48. J.C. Chen and D. Wilkinson: Phys. Rev. Let. 55, 1982 (1985)
49. F. Family, Y.C. Zhang and T. Vicsek, J. Phys. A 19, L733 (1986)
50. R. Lenormand and G. Daccord in Random Fluctuations and Pattern Growth,
eds. H.E. Stanley and N. Ostrowsky (Kluwer, Dordrecht, 1988) p. 69
51. K.J. Mal0y, J. Feder and T. J0ssang, Phys. Rev. Lett. 55, 1885 (1985)
52. H. Honjo, S. Ohta and M. Matsushita, J. Phys. Soc. Japan 55, 2487 (1986)
53. A. Hansen, E.L. Hinrichsen and S. Roux, J. Phys. A 22, L799 (1989)
54. H. J. Herrmann in Random Fluctuations and Pattern Growth, eds. H.E. Stanley
and N. Ostrowsky (Kluwer, Dordrecht, 1988) p.149
55. P. Meakin, G. Li, L.M. Sander, E. Louis and F. Guinea, J.Phys. A 22, 1393
(1989)
56. L.D. Landau and E.M. Lifshitz, Theory of Elasticity (Pergamon, Oxford, 1986)
57. S. Feng and P.N. Sen, Phys. Rev. Lett. 52, 216 (1984)
58. P.N. Sen and M. Thorpe, Phys. Rev. B 15, 4043 (1977)
59. P.G. de Gennes, J. Physique Lett. 37, LI (1976)
60. H. Yan, G. Li and L.M. Sander, Europhys. Lett. 10, 7 (1989)
61. S. Roux and E. Guyon, J. Physique Lett. 46, L999 (1985)
62. W. Nowacki, Theory of Micropolar Elasticity, (Springer, Udine, 1972)
63. E. Louis, F. Guinea and F. Flores in Fractals in Physics eds. L. Pietronero
and E. Tossatti (Elsevier, Amsterdam, 1986), p.117; E. Louis and F. Guinea,
Europhys. Lett. 3, 871 (1987)
64. H.J. Herrmann, J. Kertesz and L. de Arcangelis, Europhys. Lett. 10,147 (1989)
65. G. Peng and D. Tian (preprint)
66. G.G. Batrouni and A. Hansen, J. Stat. Phys. 52, 747 (1988)
67. E.L. Hinrichsen, A. Hansen and S. Roux, Europhys. Lett. 8, 1 (1989)
68. P. Meakin, G. Li, L.M. Sander, H. Yan, F. Guinea, O. Pia and E. Louis
(preprint);
P.Meakin in Crystal Properties and Preparation, vol.17-18, (Trans. Tech. Pub.,
Switzerland, 1989), p.l
69. H.J. Herrmann, J. Kertesz and L. de Arcangelis, (in preparation)
70. J. Fernandez, F. Guinea and E. Louis, J. Phys. A 21, L301 (1988)
71. P. Garik, R. Richter, J. Hartmann and P. Ramanlal, Phys. Rev. A 32, 3156
(1985)
72. F. Family, D.E. Platt and T. Vicsek, J. Phys. A 20, L1177 (1987)
73. M.J. Blackburn, W.H. Smyrl and J.A. Feeney, in Stress Corrosion in High
Strength Steels and in Titanium and Aluminium Alloys, B.F. Brown ed.,
(Naval Res. Lab., Washington, 1972) p.344
74. B.B. Mandelbrot, D.E. Passoja and A.J. Paullay, Nature 308, 721 (1984)
75. C.H. Scholz and C.A. Aviles, The Fractal Geometry of Faulting (1986)
290 J. Kertesz / Dielectric breakdown

76. C.S. Pande, L.E. Richards, N. Louat, B.D. Dempsey and A.J. Schwoble, Acta
Met., 35, 1633 (1987)
77. Z.Q. Mu and C.W. Lung, J. Phys. D 21, 848 (1988)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
© Elsevier Science Publishers B.V. (North-Holland), 1990 291

9. Simple kinetic models for material failure and


deformation

Paul Meakin t

9.1 Introduction
Material failure and deformation are complex phenomena that typically
involve a variety of processes occurring on a wide range of time and length scales.
For this reason both computer simulations and theoretical approaches have, so far,
been of limited value in developing a comprehensive understanding of most prac-
tically important failure and deformation processes. Considerable effort has been
devoted to the propagation of a single crack tip in a homogeneous medium. This
is an important fundamental problem, but in many cases materials are inhomoge-
neous, crack geometries are complex and crack-crack interactions are important.
In order to make progress towards the development of a better understanding of
processes such as cracking, forging and drawing, it is necessary to resort to simple
models that focus attention on those processes that are believed to be the most
important for a particular system and set of conditions. At this stage, it seems
most unlikely that tractible general models will soon be developed that can be
applied to a wide variety of materials under a broad range of conditions.
One approach that has proven to be of value is to represent the mate-
rial by a network of structural units (bonds, springs, beams, etc.) whose rate of
failure depends on the local conditions (temperature, strain, radiation, chemical
environment, etc.). In some models formation as well as failure of the elements of
the network is allowed. In most cases these basic features are incorporated into a
Monte Carlo simulation. The basic idea is to select an element of the network at
random with a probability Ρ(Ω) that depends on the local configuration (Ω). Af-
ter a bond has been selected, it is removed from the network (or its properties are
changed) and the new network is allowed to partially or completely relax to a new
mechanical equilibrium. The new network element breaking probabilities are then
* Central Research and Development Department, E. I. du Pont de Nemours and Company,
Wilmington, DE 19880-0356, USA.
292 P. Meakin / Simple kinetic models

calculated and the process described above is repeated many times to simulate the
material failure or deformation process.
It is a matter of common experience that the rate of failure of many mate-
rials increases very rapidly with the stress applied to the material and it is natural
to suppose that this global observation has a microscopic origin. Consequently, it
seems reasonable to explore the properties of models in which the failure rate R4
of network elements is related to the local force / associated with these elements
by simple relationships such as
Ri oc P (9.1)
or
R{ oc e a ( / n ) . (9.2)
In equation (9.1), a large value for the exponent η is usually appropriate. The
form given in equation (9.2) is supported by the absolute reaction rate theory for
chemical processes.^1-3^ According to this theory the rate constant fc' of a chemical
process (such as bond breaking) can be written as
fc' = 6(T)e- f i ' / f c B T (9.3)
where E& is the activation energy, fcB is the Boltzmann constant and T is the
temperature. In most cases, where ΕΆ > fcT, the temperature dependence of the
"pre-exponent factor" b(T) can be neglected. The effect of exerting an external
force on the system is to reduce the activation energies Ea by an amount that
depends on the local force / (or the local bond length displacement δ). For a
harmonic potential the elastic energy stored in the distorted bond is given by
Es = k62/2 = f2/2k (9.4)
and if it is assumed that the activation energy for bond breaking is reduced by an
amount proportional to Es, then a value for 2 is obtained for the exponent n in
equation (9.2). An alternative approach might be to assume that the activation en-
ergy is reduced by an amount proportional to the equilibrium energy difference for
the whole network before and after bond breaking. Unfortunately, this procedure
would, for most systems, require very large amounts of computer time and would
be either completely impractical or restricted to very small systems. It might also
be argued that the bond has broken when the bond length displacement δ reaches
a critical length S0 and that for a harmonic potential the energy required to reach
this displacement will be reduced by an amount proportional to / (n = 1). None
of these simple ideas is self-consistent and at this stage equation (9.2) must be
regarded as an entirely empirical relationship. For polymers there is a considerable
amount of experimental data which supports the idea that the activation energy
is reduced by an amount that is linearly proportional to the stress σ (see refer-
ences [4-7], for example). Under these conditions the bond breaking rate constant,
ϋ, can be expressed in the form

v = v0e-<E'-Wk°T, (9.5)
2. The Dobrodumov-EVyashevich model 293

where v0 is the thermal vibration frequency. For three-dimensional systems the


parameter ß in equation (9.5) may be interpreted as an activation volume. For
two-dimensional systems ß is an activation area and for the breaking of a bond in
a network the term βσ can be replaced by ß'f where ß' is an activation length.
In the limit T —> 0 the bond with the highest strain will always break if all of the
bonds are equivalent.
These models can be used to investigate both the kinetics of crack growth
and the morphological aspects of cracking processes. A time scale can be introduced
into models in which elements of the network are selected at random. After an ele-
ment has been selected, a random number X uniformly distributed over the range
0 < X < 1 is generated and the element is removed or modified if X < Pi/Pm3iX
where Pi is the removal (or modification) probability of the selected element and
Pmax is the maximum probability for any element in the system. Each time an
element in the network is selected (whether or not it is removed or modified), the
time is incremented by an amount dt given by

d* = l/(7VP max ) (9.6)

where TV is the number of unbroken (or unaltered) network elements. The time
scale introduced in this manner is directly proportional to the real (physical) time
of the cracking or deformation process.
If the bond breaking rate constants are known (for example, if the frequency
Vo in equation (9.5) is known), then an absolute time scale can be introduced in a
similar fashion by incrementing the time by an amount dt given by

d* = l / ( 7 V i w ) (9.7)

after each element in the network has been selected. Here fmax is the maximum
bond breaking rate constant. In most cases there are considerable uncertainties
concerning both the detailed physical description of material failure processes and
the parameters used in the models. Consequently, a quantitative agreement be-
tween time scales of computer models and real systems is usually either fortuitous
or a result of "tuning" the model parameters. Information concerning relative time
scales for different processes obtained from kinetic models for mechanical behaviour
is often much less sensitive to model parameters.

9.2 The Dobrodumov-El'yashevich model


A three dimensional model for brittle fracture in polymers based on equa-
tion (9.5) was developed by Dobrodumov and El'yashevich more than 15 years
agoJ8! Dobrodumov and El'yashevich used an array of (20 x 20 x 10) (x — y — z)
undeformable elements located at the sites of a cubic lattice. These elements were
joined by Hookean bonds with an elastic constant of fci in the z direction and k2
in the x and y directions. A constant external force was applied to each of the
elements located in the top and bottom (x — y) faces in the z direction. Motion
294 P. Meakin / Simple kinetic models

was restricted to the z direction and only bonds in the z direction were allowed to
break. In the Monte Carlo algorithm used in this work one of the bonds in the z
direction is selected at random with a probability Pi given by

/N υ
Ρι = νί/Σ ν (9.8)

where the summation is over all of the N non-broken bonds. The bond breaking
rate constants vi are given by
E
Vi = Voe-l --MVk*T. (9.9)

The selected bond is then broken and the system is allowed to find a new mechan-
ical equilibrium (with the motion of the undeformable elements restricted to the
±z directions). All of the bond breaking probabilities are then recalculated using
equation (9.6) and the procedure outlined above is repeated time and time again
until all of the 400 bonds in one of the (x — y) planes is broken (since only bonds
in the z direction are allowed to break, complete failure does not occur untill all
of the bonds in an (x — y) plane have been broken). Before each bond is broken,
the time is increased by an amount r* given by

' 3=1

In this model the distribution of forces fa and consequently the sequence


of events leading to failure depends only on the parameter ß1f /k^T.
The total time t = Σ T{ required for the system to fail was measured and
log(i) was found numerically to depend linearly on the external force / applied
to each of the elements on the exposed (x — y) faces. The time t was found to be
insensitive to the force constant ratio ki/k2 (over the range 1 < ki/k2 < 10) and
to increase only slightly on reducing the size of the network. At low values for the
external force / (β'f < knT) the location of the broken bonds in the early stages
of failure is quite random. For larger values of / (ß'f > kBT) stress concentration
effects are more important and crack-like patterns appear at an earlier stage in the
simulations.
A two-dimensional model closely related to that of Dobrodumov and
El'yashevich was developed by Termonia and Meakin.^ The simulations were car-
ried out using a 100 x 60 (x — y) node network with periodic boundary conditions
in the lateral x direction. In this model the bond breaking probabilities are given
by
p. K e ^i,V2fc B r ) (9 n )

where k, is the force constant associated with the ith bond (kx or ky) and Si is
the elongation. In this model the system is stretched along the y direction and
2. The Dobrodumov-EVyashevich model 295

Figure 9.1 Results from a typical simulation carried out using the two dimensional model of
Termonia and MeakinJ9^ At the start of this simulation the network of nodes and bonds was
stretched by 20% in the y direction (without allowing bond breaking to occur). The parameters
used to obtain these results were kx = 20, ky — 20 and k&T = 1. Figures a, b and c show three
different stages well before failure, close to failure and at failure. The small fragments are a
consequence fact that the network is not relaxed after each bond breaking event.

the positions of the nodes at the two ends (y = 0 and y — t/max) are fixed at the
beginning of each simulation. Figure 9.1 shows a typical result obtained from this
model. Despite the use of overrelaxation and block relaxation^ 6-8 hours of CPU
time on an IBM 3081 D computer were required. The results shown in figure 9.1
indicate that the cracks formed by this model have very irregular shapes and that
small cracks quite strongly resemble larger cracks.
This suggests that the cracks might have a fractal^11! geometry. The effective
fractal dimensionality D of these cracks was measured from the dependence of the
number of line segments n(l) of length / needed to "cover" the perimeter as a
function of /. For a self-similar fractal n(l) is related to / by

n(l) ~ ΓΌ. (9.12)

A value of 1.27±0.02 was obtained from D using lengths / in the range 2 < / <
30 in units of the equilibrium distance between the network nodes. A value of
1.31±0.15 was,obtained from the dependence of the crack perimeter P(l) on the
crack area A(l) using line segments of a fixed length / to measure the perimeter.
The dependence of P(l) on A(l) for a compact two-dimensional object with a self-
similar boundary is given by
P(l) ~ A(l)D/\ (9.13)
This result was obtained using several different values for / in the range 1 < / < 8.
The values found for D were insensitive to the model parameters though it is
clear that quite different results would be obtained for T —> 0 (D —> 1). In the
296 P. Meakin / Simple kinetic models

limit T —> oo the bond breaking process would be random and the methods used
here to measure D should measure the hull of a percolation cluster which has
a fractal dimensionality of 7/4J 12-15 ! Since the fracture process is anisotropic, it
seems quite likely that the cracks have a self-affine rather than a self-similar fractal
geometry. It would be interesting to carry out larger scale simulations in order to
determine if the cracking patterns generated by this model are self-affine or self-
similar. Unfortunately, this would probably be beyond current computer resources.

9.3 Surface cracking models


Thin film deposits have a wide range of important technical applications as
a result of their unique physical and chemical properties. In most of these systems
the surface layer and substrate have quite large differences in physical and chemical
properties. A consequence of this is that large stresses frequently occur in the sur-
face layer leading to the formation of characteristic cracking patterns that resemble
mud cracking in dried-up lake beds. A few examples of systems in which this type of
cracking occurs include cesium fluoride CeF 3 vapor deposited onto germanium,^
alumina AI2O3 sputter deposited onto copper J17^ barium fluoride deposited at low
pressures onto indium phosphide^18! and chromium metal electrodeposited onto an
aluminum alloyJ19^ The publications describing these systems were all submitted
to the same journal (Thin Solid Films) during the period March-November 1983.
This demonstrates that this type of cracking process is quite common as well as
being of scientific and technological importance. In most cases the surface layer is
under tension.
Figure 9.2 represents a model that was developed^ to simulate this type
of process. The surface film is represented by a network of bonds and nodes that
form a triangular lattice at the start of a simulation. The elastic energy associated
with this layer is given by

25 = Σ > - = Σ**?/2 (9-14)


i i

and it is assumed that the force constant k is the same for all bonds in the network.
In addition, each node in the network is attached to the substrate by a weak bond
(represented by the dashed lines in figure 9.2). The force exerted by this weak bond
on the ith node is given by
f=fc 2 (r?-r i ), (9.15)
where r^ is the position in the ith. node and r? is the position of attachment to
the substrate (its position at the start of the simulation). In this model, only
the "strong" bonds in the surface layer are allowed to break (all of the nodes
remain attached to the substrate). The bond breaking probabilities in this model
are given by equation (9.11). The simulation is quite similar to those described in
the previous section and proceeds via a series of bond breaking and relaxation steps.
Besides using block relaxation and over-relaxation, additional relaxation cycles are
3. Surface cracking models 297

Figure 9.2 A schematic representation of a model for elastic fracture in thin films. The nodes
(large dots) are connected by strong bonds to form a triangular lattice. Each node is joined to the
underlying structure by a weak bond ( ) at its original position at the start of the simulation.
Throughout the simulation the distance from the nodes to the underlying substrate is constant
(only horizontal motion is allowed). This figure shows the original configuration in which each
node is associated with six bonds in the surface network.

carried out for the positions of those nodes in the vicinity of the last broken bonds.
Periodic boundary conditions are used in both the x and y directions. In a typical
simulation^ the initial bond extension is about 10% (/ = 1.0 at the start of the
simulation and Z0 = 0.90909 or Z//0 = 1.1). Here I is the bond length and Z0 is the
equilibrium bond length. Figure 9.3 shows the results of a small scale simulation
carried out using a triangular network consisting of a 50 x 50 array of nodes with
7500 connecting bonds. Values of 1600 and 16 were used for k and fc2 respectively
and the initial bond extension (6 = l — lo) was 0.0909.... Figure 9.3b shows features
(labeled "S") which resemble shear bands.
Figure 9.4 shows results obtained from much larger scale simulations using
a network of 200 x 200 nodes joined by 120,000 bonds at the start of the simulations.
In this figure, only the locations of the broken bonds in the original (undistorted)
network are shown. All four parts of figure 9.4 were obtained using a value of 100 for
k/k2. The effects of increasing k from 600 to 2400 are shown. For large values of fc,
the first cracks to be formed are quite linear; but as the cracking process continues,
the cracks become more irregular. For smaller values of k the first cracks are less
linear and many more isolated defects are formed before large cracks appear.
This model has been used to illustrate and investigate several aspects of
surface cracking. For example, figure 9.5 shows the results of simulations similar
to those shown in figure 9.4 with a value of 800 for k and two different values for
298 P. Meakin / Simple kinetic models

b. 400 BONDS BROKEN

c. 800 BONDS BROKEN

Figure 9.3 Results obtained from a small-scale simulation (2500 nodes) with k = 1600 and
&2 = 16. The system is shown after (a) 200, (b) 400, and (c) 800 bonds broken. In (b), three of
the features that resemble shear bands are indicated with the letter S.
3. Surface cracking models 299

5000 BONDS BROKEN 5000 BONDS BROKEN

I = 200 k = 600 k2 = I = 200 k = 1000 k2 = 10

5000 BONDS BROKEN 5000 BONDS BROKEN


/ ' \ A / /

I = 200 k = 1600 k2 = 16 I = 200 k = 2400 k2 = 24

Figure 9.4 Simulations carried out using the surface cracking model with k/k2 = 100 and four
different values for k (600, 1000, 1600 and 2400 for figures a, b, c and d, respectively). These
simulations were carried out using a 200 x 200 node triangular lattice with Z0 = 0.90909 and an
initial bond length of / = 1.0.
300 P. Meakin / Simple kinetic models

2000 BONDS BROKEN 2000 BONDS BROKEN


l ," _ "* * , » ," * ,T
~ " ' ,*.,,.- · . · y- - · -..
:
*"' · ' ; V * v *.'· *" . '·, *. ;«' -' --'-:·._. Λ''-
*. ' · -' - ·; <' ■ ''«. * ', /-- ' · "" ',
· . ·-t';"«·
/'.'.'. -/ '·: ' ' * ' V \ · / ' * · · « v * ■'-'- * '' '- i '
';? : ::,%^,
■ , *' . '
". J[[r )" · . ' . « ''V..
m
, 1 » ,-'', '*
■>/'',. l /
:
■ * ' -"■ ·*
'. ^'yCw'.'i'^''
. V .'*' ' . ' jf * '" - * '-~\ '* > ■
- - ♦ " * ,. .
'-' ;vJ^>
- .'."1
'··' -v',--' . : ./'·'" ': '- -*' ..--'V.„''V
K%&; -V*'-'-V*--;C:,-*-:·
V . - : ! '·" i·,- !» '■ ■
·;-. ->
,'<->'Λ
;, * v:>" .·;;;:,·;
;---■ ^": ""?;'
■^M$, i-v
., ' '
v
. '· ζ wSji» *
*n - ·
';-- *v
k=800 k2-1.6 t= 0 0 0 0 2 6 6
; : ;. * ·""*. vL - U " - : " - ' / - · , · ' ' ~ ■'

k'*800 k2«4Ö t* 0.000442

Figure 9.5 Patterns generated by the surface cracking model with a value of 800 for k. Figure a
shows the results obtained with weak attachment to the substrate k2 = 1.6 and figure b shows re-
sults obtained with strong attachment k2 = 40. In these simulations a 200 x 200 nodes triangular
lattice was used with Z0 = 0.90909 and I = 1.0 at time t = 0.

k2 (1.6 in figure 9.5a and 40 in figure 9.5b). A comparison of figures 9.5a and 9.5b
show that "weak" attachment to the substrate leads to the more rapid growth of a
few long relatively straight cracks. This is a more serious form of failure than that
found for stronger attachment to the underlying substrate (figure 9.5b). In most
models the network is relaxed completely after each element has been modified
or removed. Partial relaxation can be used to represent "viscoelastic" effects. For
example, the time scale can be subdivided into small increments St and each of the
nodes can be moved by an amount 6Si given by

6Sj = cijöt (9.16)

during each time step. Here f; is the force on the Mi node and c is a constant. The
time increment 6t is made sufficiently small that the node displacements 6SZ are
only a small fraction (typically about 0.01) of the displacements that would bring
them into a local mechanical equilibrium with their neighbours. For large values
of c, the cracking patterns resemble those obtained with complete relaxation. For
smaller values of c the stress concentration effects are reduced and the cracks are
shorter and less regular. The total rate of bond breaking is also reduced for smaller
values of c.[20]
Simulations have also been carried out in which the bond breaking rates or
probabilities are given by equation (9.5) or

p. = p0ev6^T. (9.17)

Results obtained from simulations with three different values for v (v = 50, 100
and 200 in units for which k^T = 1) are shown in figure 9.6. In these simulations
3. Surface cracking models 301

6 0 0 0 BONDS BROKEN

Figure 9.6 Typical patterns produced by a surface cracking model in which the bond breaking
activation energy is reduced by υδ (in units of k&T). Figures a, b and c show results obtained with
three different values for v (50, 100 and 200 respectively). In all three simulations k/k2 = 100
and I/IQ — 1.1 at the start of the simulations.
302 P. Meakin / Simple kinetic models

5000
4500
4000
g3500
O3000
0*2500 txlO
CO

§2000
§ 1500
1000
500 _l I I L
°0 .00001 .00002 .00003 .00004 00002 .00006 .00007 .00008
t

Figure 9.7 Time dependence of the number of broken bonds obtained from a surface cracking
model simulation with k — 1600 and hi = 16. In all of the simulations shown in figures 9.3-9.7
lo — 0.90909 and the initial bond length in the surface layer was 1.0.

the ratio k2/k was held at a value of 0.01. Since the network configuration depends
only on k2/k and not on their individual magnitudes, the results obtained from this
model depend only on the ratio k2/k. For large values of v (figure 9.6c), the cracking
patterns resemble quite strongly those shown in figure 9.4 for large values of k. For
smaller values of v, the effects of stress concentration on the bond breaking rates
are smaller and the cracking patterns resemble those obtained from the surface
cracking model in which the reduction of the bond breaking activation energy is
given by k62/2 with relatively small values of k. For simulations carried out with
large values of k with bond breaking activation energies reduced by kS2 (or for
large values of v), the surface cracking process proceeds in three distinct stages. In
the early stages, the rate of bond breaking is relatively slow and isolated defects are
formed. Eventually, a small crack with large stress concentrations is "nucleated"
and this crack propagates rapidly to form a long straight crack. Eventually, the
cracking process reduces the strain associated with the surface layer. The cracking
process becomes slower and the cracks generated at this stage are more irregular.
This sequence of events is illustrated in figure 9.7. This figure also illustrates the
time dependent nature of these surface cracking models.

9.4 Models for the properties and failure of polymer fibers


Development of a theoretical understanding of the mechanical properties of
polymer fibers presents a major challenge. The morphology of real polymer fibers
is complex, and a variety of different processes that are not independent of each
other contribute to the deformation and eventual failure of the fiber. It is doubtful
if a general model incorporating most of what is believed to be of importance
concerning polymer fiber structure and dynamics could be developed in the near
4. Models for the properties and failure of polymer übers 303

Figure 9.8 A two-dimensional representation of the model used to simulate the mechanical
properties of polymer fibers. At the start of a simulation the nodes (that are assumed to be
incompressible) are in contact with each other along the y axis, k\ and &2 are the elastic con-
stants for the primary and secondary bonds respectively. The missing primary bonds may be a
consequence of failure or may represent a pair of contacting chain ends resulting from the finite
molecular weight.

future, and such a model would be so complex and error prone that it is unlikely to
be of much value in developing a better understanding of how polymer fiber fails.
On the other hand, models that focus attention on just one failure mechanism such
as primary (covalent) bond breakage^21,22! or slippage^23'24! are, in most cases, too
restrictive to give either realistic results or valuable insights.
Recently, Termonia et alJ 25-27 ^ have developed a variety of models related
to that of Dobrodumov and El'yashevich (see above). However, the models of
Termonia et al. include the breaking of both "primary" and "secondary" bonds
(bonds that are parallel and perpendicular to the fiber axis in the undistorted
initial state). In addition, chain end "defects" are included by removing some of
the primary bonds before the simulation begins. In a typical simulation an array
of 6 x 6 x 1000 (x — y — z) nodes is used with periodic boundary conditions in the
lateral (x and y) directions. Two-dimensional simulations with primary bonds in
the y direction and secondary bonds in the x direction were also carried out. Figure
9.8 shows a portion of an array used in one of these two-dimensional simulations
and figure 9.9 shows some typical results obtained from the two-dimensional model.
In the simulation used to obtain figure 9.9 the bond breaking probabilities for both
the primary and secondary bonds was chosen to be proportional to ea6 . However,
in most of the simulations the breaking frequencies in these models are given by

Vi = voe-^-0^k»T, (9.18)
304 P. Meakin / Simple kinetic models

Figure 9.9 Results of the simulation for a two-dimensional array of 20 x 100 nodes: case a,
150% strain; case b, 300% strain. The probabilities for breaking primary and secondary bonds
in these highly stretched samples were chosen equal to exp(e2) for simplicity. Periodic boundary
conditions were imposed in the x direction.

where the activation energies El& and activation volumes ßi are, in general, different
for the primary and secondary bonds. The vibration frequencies v0 were assumed
to be the same (v0 = 1011 s - 1 at T = 300 itf28') for both types of bonds. Any differ-
ence between the vibration frequencies of primary and secondary bonds should be
relatively small and could be accounted for (at a fixed temperature) by modifying
the activation energies Ela. In this model for polymer fibers, the primary bonds are
assumed to correspond to the strong covalent bonds along the fiber axis and the
secondary bond represents much weaker van der Waals forces and hydrogen bonds.

This work was carried out in order to develop a better understanding


of the failure and optimum properties of high strength fibers. Consequently, the
parameters used in the three-dimensional model were selected to correspond to
what is probably the most extensively studied system of this type (highly oriented
polyethylene) and were expressed in absolute rather than reduced units. For the
primary bonds the parameters fcx = 300 GPaJ 29 ! E 1 = 25 kcal/moW28'30} and
ßi = 1.54Ä . For the secondary bonds the parameters k2 = 3 GPa, El = 0.65 kcal
mole - 1 and ß2 = 2.5 A . The value used for ΕιΛ is smaller than that normally
associated with the fracture of carbon-carbon bonds but is based on previous work
on polymersJ28'30! One of the properties of primary interest in this work was the
tensile strength of the fibers. For simulations carried out using El = 80 kcal
mole - 1 and ß = 1.54 A for the primary bonds the results were similar to those
obtained with the other parameter sets for low molecular weights M where sec-
ondary bond breaking is the major failure process. For M > 2 x 105 the parameters
El = 80 kcal mole - 1 and ß = 1.54 A give an unrealistically high value for the
tensile strength^ 31-33 ! of the fiber ( « 138 GPa). Like the other models described
here, this model is a fully time dependent kinetic model. An absolute time scale is
used so that a more meaningful comparison with experimental results can be made.
This time scale is sensitive to the model parameters and the absolute time scale
4. Models for the properties and failure of polymer übers 305

Figure 9.10 Stress-strain curves calculated for several simulated polyethylene fibers of various
(monodisperse) molecular weights. The strain rate is 100% min - 1 . Three-dimensional simple cubic
arrays of up to 6 x 6 x 1000 nodes were used in the simulations. Periodic boundary conditions
in the x direction were imposed. Curves for M — 2.2 x 104 and M = 8.2x 104 have not been
calculated in their entirety, due to a lack of computer time. The dashed line indicates a slope
equal to the theoretical modulus (= 300 GPa).

should not be taken too literally. However, this model does give valuable insight
into the effect of different relative time scales (strain rate, strain frequencies, etc.)
on mechanical behaviour. The simulations can be carried out with a constant force,
constant strain rate, periodic strain, etc. Figure 9.10 shows the results obtained
from simulations carried out at a constant strain rate (100%/min) using 5 differ-
ent molecular weights. The distribution of molecular weights was assumed to be
monodisperse and the only correlation in the defects used to represent chain ends
was that required by the selected molecular weight. For low molecular weights
(M < 8 x 104) these simulations indicate that rupture of secondary bonds (in-
termolecular slippage) is a more important process than breakage of the primary
bonds. An important feature of the model used in this work is that secondary bonds
are allowed to reform between adjacent nodes if their coordinates in the direction
of the primary bonds become equal. Primary bonds are not allowed to reform. At
high molecular weights (M > 8 x 104) the rupture of both primary and secondary
bonds is important in the fiber failure process.
Figure 9.11 shows the change in the average molecular weight as a function
of strain for two of the simulations shown in figure 9.10. In the case of the simulation
with a high (3.3 x 105) molecular weight the fracture at first seems to occur in a
306 P. Meakin / Simple kinetic models

w
■t.iv 1 1 i - 1 ι—1

K 3.105 V ^ - 3 . 3 105 "1


o
UJ
MOLECULAR

\
UJ
<
O

.L
UJ
^8.2 104
ä 10»

1 1 1 1 1 1

STRAIN %

Figure 9.11 Dependence of the average molecular weight on the strain for two of the simulations
illustrated in figure 9.10. The arrow indicates the maximum in the corresponding stress-strain
curve (see figure 9.10).

brittle manner, but the behaviour becomes more ductile as the molecular weight is
reduced. For large molecular weights the number of methylene units (with a weight
of 14 dalton) is quite large (about 2 x 104 for a molecular weight of 3 x 105) and
the behaviour of such high molecular weight systems cannot be investigated using
a model in which one node represents one methylene group. At high molecular
weights, one node was used to represent m methylene units. For ra > 25 the m
secondary bonds were broken in groups of 25 with an activation energy E\ of
25 x 0.65 kcal mole - 1 and an elastic constant k2 of 25 x 3 GPa was used for
each group of 25 secondary bonds. (The values for the frequency v0 and activation
volume ß were the same as for a single secondary bond.) For m < 25 the secondary
bonds were broken in groups of 5 units.
Figure 9.12 shows the molecular weight dependence of the maximum stress
at an elongation rate of 100%/min calculated in this way for values of m ranging
from 5 to 3200. The observation that these results can be represented quite well by a
single curve supports the use of this simple renormalization scheme. This illustrates
an important characteristic of the class of models reviewed in this chapter. The
elements of the network do not, necessarily, correspond to structural components on
an atomic scale. In many cases they should be interpreted in terms of "mesoscopic"
length scales and there may be no physical components of the system that they
5. Chain slippage models 307

1 I I l| 1 1 1 I I I I H 1 1 1 I I I I l| 1 1 I I I I I II

102 103 10 4 105 I06


MOLECULAR WEIGHT

Figure 9.12 Dependence of the strength (maxima of the stress-strain curves such as those shown
in figure 9.10) on the molecular weight for simulated monodisperse fibers. The symbols are for
different values for the number m of methyl groups per node, (v) ra = 5; ( · ) m = 25; (o) m = 100;
( x ) m = 200; (■) m = 400; ( Δ ) m = 800; (D) m = 1600; (A) m = 3200.

are intended to represent corresponding to the elements (bonds and nodes) in the
network.
All of the simulations described here are carried out assuming isothermal
conditions. This is appropriate for very small systems but in many systems tem-
perature increases play an important role in material failure and deformation. It
should be relatively simple to include energy dissipation and thermal diffusion in
these models.

9.5 Chain slippage models


Most models for material failure and deformation focus attention on the
breaking of primary (covalent or metallic) bonds. However, in polymers chain slip-
page and the release of entanglements can play an important, sometimes dominant,
role. Termonia and S m i t h ^ have explored the properties of a polymer deforma-
tion model that assumes that chain slippage through entanglements and the effect
of weak van der Waals interactions on polymer motion plays the dominant role in
tensile deformation. In this model both the rate of failure of van der Waals "bonds"
and the rate of chain slippage are given by equation (9.5) with different values for
the activation energies ΕΆ and activation volumes β. For the chain slippage process
308 P. Meakin / Simple kinetic models

X * (a) X (b)

Figure 9.13 (a) Schematic representation of an undeformed polymer solid. Prior to deforma-
tion, the chain segments between entanglements are connected through numerous weak inter-
and intramolecular bonds, e.g., van der Waals forces (dotted lines). The latter provide the ini-
tial stiffness to the material. Since the coordination number of an entanglement is only 4, the
three-dimensional network has been represented by a planar network, (b) A more schematic rep-
resentation of the network, in which the details of the chain configurations are omitted altogether.
The individual van der Waals bonds are replaced by "effective" bonds (dotted lines) joining each
entanglement point to its neighbours. The heavy solid lines denote chain vectors between entan-
glements; chain ends are indicated by dashed lines. The chain vectors in an actual, undeformed
specimen are randomly oriented in three-dimensional space; i.e. (cos 2 (0)) = 1/3, where Θ is the
vector's angle with the draw axis. In the simplified, regular two-dimensional representation (b),
the same initial value Θ = 54.7° was taken for all vector orientations along the y axis.

the parameter σ in equation (9.5) represents the stress difference between two parts
of a chain separated by an entanglement point.
The two-dimensional model used in this work is illustrated in figure 9.13.
At the start of a simulation the system consists of a two dimensional "diamond"
lattice of nodes representing the entanglement points between pairs of polymer
molecules. The lattice is then decorated randomly with polymer molecules that
intersect only at the nodes (entanglement points) until there is an entanglement
point associated with every node. The stress σ in the portions of the polymer
chains separating entanglement points is given by the classical theory of rubber
elasticity, t35^
σ = akBTC~l (—\ . (9.19)
5. Chain slippage models 309

Here n is the number of statistical chain segments of length I between a pair of


entanglement points separated by a distance r and C~l is the inverse Langevin
function (C(x) = coth(:r) — 1/x). The parameter a is given byt35l

a = (Nß)y/n, (9.20)

where TV is the number of chain strands per unit volume.


Like the other models discussed here, the simulation consists of a sequence
of bond breaking and relaxation steps. At frequent intervals the network is also
strained along the y axis. In this model only displacements in the y direction are
calculated explicitly. The coordinates in the x direction are assumed to be reduced
uniformly by 1/y/X where λ is the draw ratio (l/lo) and periodic boundary condi-
tions are used in the x direction (figure 9.13). Perhaps the most drastic assumption
made in this model is that once van der Waals bonds have been broken they are
not allowed to reform. A discussion of the selection of parameters (activation ener-
gies, activation volumes, etc.) used in this model and their justification is beyond
the scope of a general survey such as this. However, two important parameters are
the molecular weight M 0 between entanglement points and the polymer molecular
weight M. Figure 9.14 shows the patterns generated by a simulation carried out
with a molecular weight of 1900 between entanglement points (a value character-
istic of molten polyethylene). At low molecular weights M « M 0 brittle fracture
is observed (fig. 9.14a). At M = 5M 0 (fig. 9.14b) a "neck" forms and propagates
along the sample. For M ~ 10.5M0 (fig. 9.14c) the deformation is more homoge-
neously distributed throughout the sample and for very large molecular weights
(fig. 9.14d) the deformation is quite homeogeneous, at least for strain ratios up to
λ = 3. In qualitative terms these results are in good agreement with experimental
observations J4'5'29'35'36]
A more serious attempt to compare simulation results with experiments
was carried out later by Termonia and SmithJ37^ Figure 9.15 shows the results
of simulations similar to those shown in figure 9.14. However, in this case the
(monodisperse) molecular weight is high in all cases (corresponding to a molecu-
lar weight of 475,000 for polyethylene) and the quantity that is varied is M0 the
molecular weight between entanglements. This quantity is expressed in terms of
the ratio φ given by
φ = 1900/Mo (9.21)
where 1900 is the value for M 0 in a polyethylene melt. Brittle fracture (fig. 9.15a) is
again observed when M 0 « M (the molecular weight of the polymer). Figure 9.16
shows the results of experimental studies of the drawing of polyethylene films.
The films were prepared by solidification from the melt (M 0 « 1900) and
from decalin solutions using molecular weights of M n = 2 x 105 and M w = 1.5 x 106
respectively where Mn is the number average molecular weight and M w is the weight
average molecular weight. The qualitative agreement between figures 9.15 and 9.16
is quite striking. A more quantitative comparison between these simulations and
310 P. Meakin / Simple kinetic models

Figure 9.14 Typical "morphologies" obtained with a model for polyethylene of different monodis-
perse molecular weights: (a) M = 1900; λ = 1.75; (b) M = 9500, λ = 2; (c) M = 20 000, λ = 2.5;
(d) M = 250 000, λ = 3. Here λ is the draw ratio.
5. Chain slippage models 311

Figure 9.15 Simulated morphologies obtained from a model for the deformation of semicrys-
talline polymers. The parameters used in this model were selected to correspond to polyethylene
with different values for the parameter φ = MQ/M. (φ = 0.004, 0.02, 0.1 and 1 in figures a, b,
c and d respectively). In cases b - d the draw ratio λ = 2.7. The widths of the "samples" are
respectively 3.2, 1.6, 0.7, and 0.2 mm.
312 P. Meakin / Simple kinetic models

Figure 9.16 Micrographs of drawn samples of polyethylene films of Mw = 1.5 x 106 and
Mn ^ 2 x 105 crystallized from solutions in Decalin and from the melt (see reference [37] for
experimental details); these samples were drawn to a macroscopic draw ratio of approximately
3 at 100°C. The initial polymer volume fractions were respectively (a) φ = 0.005, (b) φ — 0.02,
(c) φ = 0.1, (d) φ = 1. Parts a, b and d are optical micrographs taken under crossed polarizers
(except b), and c is a scaling electron micrograph. The width of all strips shown is 0.1 mm.
6. Analog models 313

experiments is shown in figure 9.17. For very small values of φ (Μ « M0) stress
concentration resulting from failure of the van der Waals bonds is not distributed
by chain entanglements. Consequently, failure is highly localized and occurs at
quite small strains. The agreement between figures 9.17a and 9.17b is not perfect,
but in view of the simplicity of the models, it is quite gratifying. It should be noted
that in this model the "van der Waals bonds" should not be interpreted in terms
of individual van der Waals interactions. Instead, the van der Waals bonds have
an activation energy Ea of 30 kcal mole - 1 and the irreversible fracture of one of
these bonds represents the failure of a large number of van der Waals interactions.
This process can be regarded as an irreversible change in the material structure on
a nm length scale which may indeed occur on the time scale of the experiments.
Essentially the same model has been applied by Termonia and W a l s h ^
to the mechanical properties of glassy polymers. Results for the dependence of
the yield stress on the strain rate were compared with the experimental results of
Haward^40! for PMMA (poly(methyl methacrylate)). The simulation results at two
different temperatures are in quite good agreement with the experimental data.
Figure 9.18 shows the change in the deformation morphology as the density of
entanglements is increased. As the entanglement density is increased, a transition
from a crack-like pattern to the formation of a more uniform yielding zone can be
seen.

9.6 Analog models


While analog devices have been used for many years to explore a broad
range of phenomena, this approach has fallen into disfavor because of the difficulty
of precisely constructing and controlling them and because of the rapidly increas-
ing capability and ease of use of digital computers. In recent years, the availability
of large numbers of uniformly sized and shaped polystyrene microspheres with
diameters in the 1-10 mm r a n g e ^ has made possible a variety of analog simu-
lations that provide an important bridge between experiments and simulations.
While precise control and measurement of such particles is still difficult, they can
be used to explore larger system sizes and longer times than is possible with digital
simulations.
One application of these systems that has been pioneered by Skjeltorp^42,43^
is the study of cracking in polystyrene bead monolayers. The monolayer is con-
structed by placing an aqueous dispersion of the microspheres between two parallel
sheets of glass and allowing the water to evaporate slowly along one edge of the gap
between the two glass sheets. The separation between the glass sheets is maintained
by a small fraction of particles that are slightly larger than the otherwise almost
monodisperse distribution. A typical system prepared in this manner consists of a
polycrystalline monolayer with approximately 107 particles and a mean grain size
of 105 — 106 particles. The grains consist of almost perfectly close packed spheres
with a small number of isolated defects (mainly particle vacancies). In a typical
314 P. Meakin / Simple kinetic models

20 30 40
STRAIN

1 —ι 1 1 i T 1 I

30

u

2

i3 20 H

"
NOMINAL

0.1
3

O02
0.011
KöjöoTi 1 1 1 1 _J 1
20 40 60
STRAIN

Figure 9.17 a: Calculated nominal stress-strain curves for monodisperse linear polyethylene with
M = 475,000 at four different values, indicated in the graph, of the entanglement spacing factor
φ (see text). The deformation temperature was 100°C, and the rate of elongation was 500%/min;
b: Experimental stress-strain curves recorded at 120° C for ultrahigh molecular weight (UHMW)
polyethylene films M w = 1.5 x 10 6 , M n « 2 x 10 5 ) crystallized from the melt and from solutions
of various initial polymer concentrations: 10, 2, 1 and 0.5% by volume of the polymer. Prior
to deformation, the solvent had been removed from these films to reveal only the effect of the
initial polymer volume fraction and to eliminate possible plasticizing effects of the solvent (see
reference [7] for further experimental details).
6. Analog models 315

Figure 9.18 Deformation patterns obtained for notched "PMMA" samples in tension
M = 165,000. (a) Before deformation; (b) "PMMA" at 15% external strain; (c) same specimen
as in (b) but at 22.5% external strain, (d) polymer glass at 15% external strain and a density of
entanglements three times higher than in (b) but at 22.5% external strain. Temperature was set
equal to 22°C and rate of deformation e = 0.01 s _ 1 . The network has 50 nodes along the tensile
y axis and 200 nodes in the average distance between entanglements is of the order of 1 mm. The
dark lines within the deformed regions in (b) to (e) denote extended chain strands.
316 P. Meakin / Simple kinetic models

Figure 9.19 Optical micrographs of cracking patterns in a single grain of a polycrystalline


monolog of sulfonated polystyrene microsphere confined between two parallel sheets of glass.
Figure a shows an early stage in the crack growth process and figure b shows a much later stage.

Figure 9.20 Simulations carried out using the modified surface cracking model with a network
of 40,000 nodes and 120,000 bonds. Figure a was obtained using the parameters k = 400, k2 = 40
and 6m&x = 0.25. The initial strain (σ = h/lo) was 1.2, that corresponds quite closely to that
used in the experiment (δι = 3.4 mm, δο = 2.7 mm, δ\/6ο = 1.26); 3,000 bonds have been broken
at the stage shown in this figure. Figure b was obtained using the parameters k = 300, k2 = 30,
<$max = 0-15 and 1\/IQ = 1.2. 7500 bonds have been broken at the stage shown here.

cracking simulation a monolayer of 3.4 μιη (±1%) spheres is generated at the be-
ginning and the monolayer is allowed to dry slowly. During this drying process the
particle diameter of the sulfonated polystyrene microspheres decreases to 2.7 μτα.
This produces a strained film that cracks in a fashion similar to that associated
with the surface cracking model described above.
Figure 9.19 shows an early stage in the growth of a cracking pattern inside
7. Summary 317

one of the grains (cracking first occurs at the grain boundaries). The cracking
within a grain follows the same sequence of events as was found using the surface
cracking model described earlier. At first, small isolated defects are formed. Then
a rapid growth or more or less linear crack (figure 9.19a) is observed. At the later
stages the strain in the monolayer is reduced and the cracking process becomes
slower. At the same time the cracks become more irregular (figure 9.19b).
In order to represent the processes occurring in this analog model more
accurately, the surface cracking model described above was modified to allow the
points of attachment of the network nodes to the substrate to move. If the dis-
tance |r? — Ti\ (equation (9.15)) increases to a value greater than £ max , the point
of attachment to the substrate at position r? is moved towards the current posi-
tion of the node r^ until |r° — r^l becomes equal to 6mSLX. Figure 9.20 shows some
results obtained from simulations carried out using this model. Figure 9.20a shows
a relatively early stage in a simulated cracking process. The pattern generated by
this simulation resembles quite closely that shown in figure 9.19a. Figure 9.20b
shows a later stage in a cracking pattern obtained using a somewhat different set
of parameters. In this case the cracking pattern looks more like that shown in
figure 9.19b.
We do not know if the cracking of the microsphere monolayer can be well
represented by the surface cracking model described above or if fracture proceeds
via thermally activated bond breaking events in which the reduction in activation
energy is proportional to kS2/2. However, it is apparent that the analog and digital
models have a number of important features in common. In both cases, a thin layer
of material is connected to an underlying essentially rigid substrate by relatively
weak forces (in the analog experiment the microsphere monolayer may be attached
to both of the treated glass surfaces but this detail is not of major concern here).
The rate of bond breaking or crack propagation is sensitive to the strain in the
surface layer. The systems evolve from a well ordered initial state with a relatively
simple stress/strain field to a disorderly system with a complicated stress/strain
field. These features seem to be the essential ingredients in the sequence of events
seen in both the analog and digital models. The strain history is different in the
analog and digital models but this is not believed to be very important since in-
tragranular failure in the analog model begins only after the strain has approached
its maximum value.

9.7 Summary
I have attempted to illustrate in this review how quite simple models can be
used to illustrate and understand failure and deformation processes in real materi-
als. Simple models are also used to explore the universality and scaling properties
of mechanical systems (see chapters 5-8 in this volume). Since these properties
are associated with the asymptotic behaviour of mechanical systems, a reliable
assessment requires large scale simulations (and many of them are required to re-
318 P. Meakin / Simple kinetic models

duce statistical uncertainties). At present, the numerical methods used to relax


mechanical networks present a serious barrier to further progress in this direction.
The applications emphasized in this chapter also require substantial computer re-
sources, but in this case the limitations associated with present day algorithms
and computers are not as severe. Despite the pioneering work of Dobrodumov and
EryashevishJ 8 ] there has still been relatively little work carried out with three-
dimensional models. This is surprising since the demand on computer resources is
not inordinately more severe than that required for two-dimensional simulations
and the algorithms would not be much more complex. It is apparent from the work
described in this short survey that even quite simple models can provide valuable
insight into the deformation and failure of real materials. These simple models pro-
vide a sound basis for the development of more realistic and more complex models
that should better describe the behaviour of real materials. The judicious inclusion
of additional processes in models of the type described here should prove to be
of value in obtaining a better description of the mechanical properties of real ma-
terials that may eventually be of predictive value in the design of new materials.
At the same time, improved computer resources and algorithms should allow us to
develop a better understanding of the fundamental aspects of material deformation
and failure using simple models like those described here. The development of a
comprehensive understanding of the mechanical properties of materials presents a
serious challenge to simulators, experimentalists and theoreticians alike.
My interest in models for material failure has been stimulated by active
collaborations with a number of colleagues including Francisco Guinea, Gang Li,
Enrique Louis, Oscar Pia, Leonard M. Sander, Paul Smith and Hong Yan. The work
on surface cracking models was carried out with Arne T. Skjeltorp who provided
figure 9.19. I would particularly like to acknowledge the contributions made by
Yves Termonia to this chapter. To a large extent this chapter consists of a review
of some of his recent research contributions. Figures 9.1 and 9.9-9.18 are taken
from his work.

References
1. H. Eyring, J. Chem. Phys. 3,107 (1935)
2. W.F.K. Wynne-Jones and H. Eyring, J. Chem. Phys. 3, 492 (1935)
3. S. Glastone, K. Laidler and H. Eyring, Theory of Rate Processes, (McGraw
Hill, New York, 1959)
4. A. J. Kinloch and R. J. Young, Fracture Behavior of Polymers, (Applied Science,
London,1983)
5. H.H. Kausch, Polymer Fracture, (Springer-Verlag, Berlin, 1987)
6. S.N. Zhurkov and E.E. Tomashevsky in Proceedings of the Conference on the
Physical Basis of Yield and Fracture, Oxford, (1960), p. 200
7. S.N. Zhurkov, V.S. Kuksenko and A.I. Stutsker in Proceedings of the Second
International Conference on Fracture, Brighton (1969), p. 531
References 319

8. A.V. Dobrodumov and A.M. EPyashevich, Sov. Phys. Solid State 15, 1259
(1973). (Fiz. Tverd Tela 15, 1891 (1973))
9. Y. Termonia and P. Meakin, Nature 320, 429 (1986)
10. D.N. de G. Allen Relaxation Methods, (McGraw-Hill, New York 1954)
11. B.B. Mandelbrot, The Fractal Geometry of Nature, (W.H. Freeman and Com-
pany, New York, 1982)
12. R.F. Voss, J. Phys. A 17, L373 (1984)
13. R.M. Ziff, Phys. Rev. Lett. 56, 545, (1986)
14. H. Saleur and B. Duplantier, Phys. Rev. Lett. 58, 2325 (1987)
15. A. Coniglio, N. Jan, I. Majid and H.E. Stanley, Phys. Rev. B35, 3617 (1987)
16. S. F. Pellicori, Thin Solid Films, 113, 287 (1984)
17. R. Jarvinen, T. Mantyla and P. Kettunen, Thin Solid Films, 114, 311 (1984)
18. J.M. Phillips, L.C. Feldman, J.M. Gibson and M.L. McDonald, Thin Solid
Films, 107, 217 (1983)
19. E. Namgoong and J.S. Chun, Thin Solid Films, 120, 153 (1984)
20. P. Meakin, Thin Solid Films, 151, 165 (1987)
21. S.N. Zhurkov and T.P. Sanhirova, Dolk Akad. Nauk SSSR 101, 237 (1955)
22. F. Bueche, J. Appl. Phys. 26, 1133 (1955); 28, 784 (1957); 29, 1231, (1958)
23. A. Tobolsky and H. Eyring, J. Chem. Phys. 11, 125 (1943)
24. J.R. Schaefgen, T.I. Bair, J.W. Ballou, S.L. Kwokck, P.W. Morgan, M. Panar
and J. Zimmerman in Ultra-High Modulus Polymers, A. Ciferri and I. M.
Ward, eds., p. 199, (Applied Science Publishers, London, 1979)
25. Y. Termonia, P. Meakin and P. Smith, Macromolecules 18, 2246 (1985)
26. Y. Termonia, P. Meakin and P. Smith, Macromolecules 19, 154 (1986)
27. Y. Termonia, S. R. Allen and P. Smith, Polymer 27, 1845 (1986)
28. S.N. Zhurkov, V.l. Vettegren, V.E. Korsukov and I.I. Novak Proceedings of the
2nd International Conference on Fracture, (Chapman and Hall Ltd. London,
p. 545, 1969)
29. I.M. Ward, Mechanical Properties of Solid Polymers, 2nd edition, p. 270, (Wi-
ley, New York, 1983)
30. S.N. Zhurkov and V.E. Konsukov, J. Polymer Sci-Polymer Phys. ed, 12, 385
(1974). Sov. Phys.-Solid State (Engl. Trans) 15, 1379 (1974)
31. K.E. Perepelkin, Angew Makromol. Chem. 22, 181 (1972).
32. D.P. Boudreaux, J. Polymer Sei., Polym. Phys. Ed. 11, 1285 (1973)
33. B. Crist, M.A. Rafner, A.J. Brower and J.R. Sabin, J. Appl. Phys. 50, 6047
(1979)
34. Y. Termonia and P. Smith, Macromolecules, 20, 835 (1987)
35. L.R.G. Treloar, The Physics of Ruber Elasticity, 2nd ed., (Clarendon, Oxford,
1969).
36. J.D. Ferry, Viscoelastic Properties of Polymers, 3rd ed., (Wiley, New York,
1980)
37. Y. Termonia and P. Smith, Macromolecules, 21, 2184 (1988)
38. P. Smith, P.J. Lemstra, J.P.L. Pijpers and A.M. Kiel, Colloid Polym. Sei. 259,
320 P. Meakin / Simple kinetic models

1070 (1981)
39. Y. Termonia and D.J. Walsh, J. Mater. Sei. 24, 247 (1989)
40. R.N. Haward,Tie Physics of Glassy Polymers, (Applied Science, London, 1973)
41. J. Ugelstad, P.C. Mork, K.H. Kaggerud, T. EUingsen and A. Berge, Adv. Col-
loid Interface Sei 13, 101 (1980) (manufactured by Dyno Particles A/S, P. O.
Box 160, N-2001, Lillstr0m, Norway)
42. A.T. Skjeltorp in Time Dependent Effects in Disordered Materials, NATO ASI
Series B 167, R. Pynn and T. Riste eds., (Plenum Press, New York, 1987) p. 1
43. A.T. Skjeltorp and P. Meakin, Nature 335, 424 (1988)
STATISTICAL MODELS FOR THE FRACTURE OF DISORDERED MEDIA
H.J. Herrmann and S. Roux (editors)
© Elsevier Science Publishers B.V. (North-Holland), 1990 321

10. Fragmentation

Sidney Redner*

10.1 Introduction
The breaking up of particulate material into smaller fragments is a ubiqui-
tous process that underlies many natural phenomena and technological processes.
At the geological time scale, fragmentation is responsible for the distribution of
fragment sizes that appear on beaches, on mountain boulder fields, and in soils J1-4^
Related degradation processes are responsible for the size distribution of lunar ge-
ological features^5-7] and the size distributions of asteroidsJ8~10l More practical
examples of fragmentation occur in mineral processing,^ 1_151 where the size reduc-
tion of raw ore is the initial step that must be performed in order to extract the
valuable minerals. This breakage process is energy intensive and generally quite
inefficient. Owing to the non-trivial fraction of all energy use that goes into min-
eral processing, considerable effort has been devoted to the optimal design of ore
processing machinery.
In addition to the obvious realizations of fragmentation in rock fracture,
there are many other situations where various forms of size reduction and frag-
mentation occur. At a molecular level, the breaking of individual chemical bonds
underlies polymer degradation phenomena.f16-22! In combustion and in other types
of dissolution phenomena, as the surface of the consumed object recedes, irregu-
larities form which ultimately lead to the breaking up of the materialJ 23-27 ! This
process is responsible for the size distribution of the residual ash, and an under-
standing of its properties is important in pollution control. The breakup of liquid
droplets by agitation^ 28-31 ] is an example where surface tension needs to be sur-
mounted to cause fragmentation, and the resulting aerosol formation has a myriad
of practical applications. Within the context of fluids, the breakup of eddies and
vortices in turbulent fluid flows may also be viewed as fragmentation processes,^
where the entity undergoing breakup is the flow field itself. Thus in a variety of
forms, fragmentation is the primary mechanism for many dynamic phenomena.
t Center for Polymer Studies and Department of Physics, Boston University, Boston, MA
02215, USA.
322 S. Redner / Fragmentation

In most realizations of fragmentation one is unable to observe a given


breakup event in detail, but rather, one observes the distribution of fragment sizes
that results from the breakup process. Theoretical treatments of fragmentation
have therefore been focused primarily on either predicting the fragment size dis-
tribution, based on knowing (or assuming) certain facts about the instantaneous
breakup process, or conversely, inferring details about the breakup process based
on experimental observations of the fragment size distribution. There are a wide
variety of possible instantaneous breaking mechanisms which depend on the ma-
terial properties and geometry of the object undergoing breaking, and also on the
energy input to the system. These details determine the nature of the fragment
size distribution in an instantaneous breakup event. For a continuously evolving
fragmentation process, however, particle breakup eventually takes place on all size
scales. This feature suggests the application of statistical approaches in which the
fragment size distribution from a single breakup event serves to define basic param-
eters in the dynamical rate equations that govern continuous fragmentation. The
goal of this chapter is to investigate the nature of the fragment size distributions
that arise in continuous fragmentation processes through the solution to these rate
equations.
There already exists a vast literature on fracture^33-50] and fragmenta-
51-66
tionf ! and a wide range of phenomena have been investigated, both theoreti-
cally and experimentally. Interestingly, what are regarded as immutable and basic
facts in the literature of one subfield are unknown or apparently not appreciated in
other subsets of the literature. I will therefore attempt to give a coherent presenta-
tion of some of the basic results which are representative of the broad spectrum of
physical realizations of fracture and fragmentation. In section 10.2, I first review
some basic facts about instantaneous fracture to illustrate the types of fragment
size distributions that can occur and to discuss the basic parameters that influence
the form of the distribution. The available data suggests that for a wide variety
of materials and fracture mechanisms, there is a power-law dependence of the rel-
ative concentration of fragments on their size over a non-negligible rangeJ 33 ' 40-47 ]
It is intriguing that the exponent of this power law falls within a relatively nar-
row range of values for a wide variety of systems. This distribution of fragment
sizes in instantaneous breakup serves as a basic ingredient in determining the dy-
namics of continuous fragmentation. Thus it is important to use the experimental
observations to develop mathematical models that have a realistic basis.
In section 10.3, I will elucidate some of the recent progress made in un-
derstanding the nature of the fragment size distributions that occur in continuous
fragmentation. I will focus primarily on the investigation of this problem by the
rate equationsJ 51 ' 55,58-66 ' The rate equations are an approximation of a mean-field
character, as fluctuations in the system are ignored. That is, fragments are assumed
to be distributed homogeneously at all times throughout the system, as might ac-
tually occur in a milling process, and the variability of cluster shapes is ignored.
Thus the mass is the only dynamical variable that characterizes a fragment in
1. Introduction 323

the rate equations. Although there are limitations inherent in this approach, the
rate equations do provide a comprehensive theoretical description for fragmenta-
tion which do compare well with certain experimental results. Moreover, the rate
equations may serve as the starting point for further theoretical refinements.
I will first outline some of the primary features that emerge from the exact
solutions of the rate equations for a particularly simple class of physically-motivated
models.t 58-60 ] I will then emphasize recent work in which scaling analysis is applied
to extract asymptotic solutions to the rate equationsJ63'64) This is a relatively
simple, yet powerful method which provides a universal classification scheme for the
rate equation solutions. The applicability of scaling rests on the observation that
the fragment size distribution is often characterized by a single typical cluster size,
which is a decreasing function of time. Consequently, it is reasonable to expect that
the fragment size distribution will be a function only of the ratio of the size to the
typical size. From this physical expectation, it is possible to determine the possible
asymptotic forms for the cluster size distribution quite simply. I also discuss a
special regime of behaviour, in which small clusters are more likely to break up
than large clusters, that leads to "shattering" J51'60'63,64^ This is a mathematical
pathology which describes the interesting situation where mass is lost to a shattered
phase consisting of an infinite number of zero mass particles. The general features
of the universality and scaling of the fragment size distribution and the shattering
phenomenon have parallels in the inverse process of aggregationJ 67-76 ]
In section 10.4, I give a brief summary and discuss some possible future
research directions where the current theoretical understanding is incomplete. For
example, the role of spatial inhomogeneities is largely unexplored, but clearly is a
central aspect in geological fragmentation processes. In addition, while the widely-
accepted assumption of the linearity of continuous fragmentation processes appears
to be appropriate for certain situations, linearity is clearly inappropriate in cases
where fragmentation is driven primarily by the particles in the medium rather than
by an external source. The development of models to treat the non-linear aspects
of fragmentation may be a rewarding new area of investigation.
10.1.1 Instantaneous fracture
Upon the sudden introduction of a sufficient strong external energy source,
an object can break into smaller fragments, whose distribution is the primary
observational evidence for the nature of the breakup process. For brittle materials,
breaking occurs when the applied stress exceeds the fracture stress, which is often
close to the point where linear response breaks down. According to the classical
Griffith theory, ^ the magnitude of this fracture stress depends on the distribution
of pre-existing flaws, or cracks, in the material. Owing to the concentration of stress
at a crack tip, the fracture stress of a flawed material is typically much smaller
than the stress needed to break atomic bonds. For example, for a two-dimensional
object containing a single linear crack of length L, the fracture stress σ is given by
σ — (2JE/L)1/2, where 7 is the surface free energy per unit crack surface area, and
324 S. Redner / Fragmentation

E is Young's modulus of the material. Thus in a finite volume of brittle material,


the largest crack in the object determines its fracture stress. Since the energy cost
of a crack is proportional to an area, while the strain energy is proportional to
a volume, the amount of energy available for crack formation at a fixed stress
decreases as the particle size decreases J11»78! Thus under conditions of fixed energy
density, one has the general fact that large objects are more susceptible to breaking
than small objects. This monotonic size dependence of the overall breakup rate of
an isolated object is one of the basic ingredients in the theoretical treatment of
continuous fragmentation.
Interestingly, in the crushing of rocks, the energy is typically furnished in
the form of uniaxial compression, or perhaps in the form of shear for the case of
comminution. However, under a localized compression, there is a redistribution of
stresses so that a considerable portion of an object of finite-size is under tension
in a direction perpendicular to the external compression.^ This tensile stress can
open cracks further, leading to ultimate failure. This mechanism appears to be the
normal mode of breaking in materials such as rocks.
There is considerable well-documented data for the distribution of fragment
sizes in a variety of realizations of instantaneous fracture, ranging from the over-
compression of glass spheres,t 41-43 ! to high-velocity impacts of small projectiles onto
fixed rock targetsJ 46 ' 47 ! On the basis of these experiments, the possible outcomes
for the fracture event are often classified into three broad regimes of behaviour
depending on the energy input and on the geometrical features of the experiment.
While the terminology is not universal, the three regimes may be written as:^11^
• abrasion (or surface erosion)
• cleavage (transition behaviour)
• destructive breaking (or shattering).
The corresponding particle size distributions for each of these regimes are
sketched in fig. 10.1. In abrasion, the breaking mechanism is applied specifically
to the surface, or the energy intensity is sufficiently small that only a small flecks
are removed from the surface of the object. This leads to a bimodal fragment
distribution, consisting of coarse product and fine product components. In cleavage,
the energy input is just sufficient to propagate of the order of one crack through
the sample. Typically, then, the resulting fragments are of the same order in size
as the original object. Finally, in destructive breaking, enough energy has been
imparted to the system so that many regions of the material are stressed beyond
the breaking point. This leads to a wide range of fragment sizes being produced,
which are typically all much smaller than the size of the original object. While this
classification is a useful scheme to broadly characterize instantaneous breaking, it is
highly idealized. In real breaking processes, there may be a continuum of outcomes
as a function of the energy input rather than a sharply defined transition between
two well-defined regimes.
For treating continuous fragmentation, the size distribution that arises from
instantaneous breaking is a fundamental building block upon which theories for
1. Introduction 325

*a a
abrasion
cleavage (coarse product)

cleavage 1
destructive
breaking

destructive breaking /abrasion 1


l / 7 \ ( f i n e product:
1*
M
I 1

* Ο θ oo FRAGMENT SIZE initial

Figure 10.1 Schematic representations of the different mechanisms for instantaneous particle
fracture and sketches of the corresponding fragment size distributions (adapted from ref. 11.

continuous breakup can be constructed. There are a number of equivalent forms for
expressing the size distribution, and the alternatives employed depend, to a large
degree, on the nature of the experiment and/or on convention. Experimentally, it is
often most convenient to consider M(r), the cumulative mass of fragments or linear
dimension less than r. This quantity is obtained naturally in sieving measurements.
Alternatively, this type of measurement also provides N(x), the cumulative number
of fragments with mass greater than x. Finally, many theoretical treatments often
express the size distribution in terms of c(x) dx, the concentration of fragments
with mass between x and x + ax. In the limit of small fragment masses (or radii)
and for a wide variety of physical systems, these distributions appear to be well-
described by the power-law formsJ33^

M(r) oc r a ,
Ν{χ)(χχ-{1~α/3\ (10.1)
{2 a/s
f{x)dx(xx- - \
where the exponents characterizing the latter two distributions follow by a sim-
ple change of variables from the distribution for M(r). In many experiments on
instantaneous and highly energetic breaking, exponent values between 0.5 and 1.0
have been quoted for a;J41~47,55^ For certain cases, this power law extends over two
decades of fragment sizes, in the small size limit. This power-law behaviour ap-
pears to be a universal aspect of instantaneous and destructive breaking of brittle
326 S. Redner / Fragmentation

objects.
Many of the theories that have been developed to describe the fragment
size distribution arising in instantaneous breaking are probabilistic in nature. They
are based on the idea that pre-existing flaws are distributed throughout a brittle
material. When the object is stressed far beyond its fracture point, these flaws
are activated and propagate until they reach the boundary of the material or
until another crack is encountered. This ultimately leads to the breaking off of
fragments whose size distribution is controlled by the location and geometry of the
flaws. Perhaps the most detailed investigation in this spirit is that of GilvarryJ41^
Under the assumption of a purely random distribution of flaws, the mass fraction
of fragments with linear size less than r was found to vary as

M{r) = 1 - exp(-r/fei - r2/k2 - r3/k3), (10.2a)

where the coefficients fci, &2> and £3 are proportional to the densities of linear,
area, and volume flaws, respectively. In the case where only one class of flaws is
predominant, the above distribution reduces to the exponential form

M{r) = 1 - exp{-rß/K). (10.2b)

This is also termed the Weibull^ or Rosin-Rammler^ distribution, with the


exponent ß the index of the distribution. Refinements of the Gilvarry approach
have been proposed by various authors in which the probabilistic aspects of the
original Gilvarry theory are treated more carefully. Klimpel and Austin^45! thus
derived the following form for M(r)

which gave excellent agreement with experiments over nearly the entire size range.
Here the parameters n^'s are constants which depend on the densities of edge, area,
and volume flaws, respectively.
In the small size limit, these distributions all reduce to the power law

M(r) oc r 7 , (10.2d)

with the exponent 7 = 1. (This apparently universal form is often called the
Gaudin-Schuhmann distribution in the mining engineering literatureJ 40 ' 44 !) Con-
siderable data for the instantaneous fracture of an isolated object are in good
agreement with the hypothesis that M(r) ~ r1 (fig. 10.2). Although this power
law is found to hold over a substantial range, the upper limit of linearity for the
cumulative mass is typically 0.1, or less. The full Gilvarry, or Klimpel-Austin forms
provide a better fit for this remaining mass fraction which belongs to the largest
fragments. It is worth noting that in the Gilvarry experiments, these largest frag-
ments are contiguous to the surface of the initial object. Gilvarry hypothesized
2. Continuous fragmentation 327

0.01

Figure 10.2 Fragment size distribution for 2.54 cm diameter glass spheres subjected to a com-
pressive load far beyond the fracture stress. Shown is the cumulative fraction of mass, M(r),
whose linear dimension is smaller than r (from ref. [43]). The slope of the linear portion of the
data is unity.

that these "surface" fragments may have a different distribution from that of the
smaller, interior fragments. The possibility that surface, or finite size effects play
a role in the fragment size distribution is appealing, but this aspect has not been
seriously investigated.
The experimental facts outlined above serve as a basic framework from
which one can develop theories for continuous fragmentation.

10.2 Continuous fragmentation


10.2.1 The rate equations
Having outlined some basic facts about instantaneous fracture, we now
build on these results to investigate the basic features and the results which follow
from the rate equation approach to continuous fragmentation. Implicit in this view-
point is the assumption that the form of the fragment size distribution that arises
in a single isolated breakup event continues to hold in continuously-evolving frag-
mentation. For this dynamic process, we now denote by c(x,t) the concentration
of clusters of mass x at time t. Notice that the concentration is taken to be spa-
tially uniform, an assumption which might be realized in a strongly mixing milling
process. Furthermore, by keeping only the cluster mass as the basic variable, one
is ignoring potential effects induced by the variability in fragment shapes.
According to these conditions, the evolution of c(x,t) may be described by
the linear integro-differential equation^58-64]
oo
dc(x,t)
-a(x) c{x, t) + J c(y, t) a(y) f(x\y) Ay. (10.3a)
at
328 S. Redner / Fragmentation

Here α(χ) is the overall rate at which x breaks, i.e., a(x)dt is proportional to the
probability that a cluster of mass x breaks in a time interval d£, while f(x\y) is
the relative breakup rate, i.e., the conditional probability at which x is produced
from the breakup of y. The first term on the right hand side of eq. (10.3a) therefore
accounts for the loss of clusters of mass x due to their breakup, while the second
term accounts for the gain of ^-clusters by the breakup of particles with mass larger
than x.
In the case of purely binary breakup, where each breakup event produces
exactly two fragments, the rate equations may be written in the alternative and
convenient form,

^ ^ = - c ( x , t) J F(y, x-y)dy + 2J c{y, t) F(x, y - x) Ay. (10.3b)


0 x

Here F(x,y) is the rate at which a cluster of mass x + y breaks up into two
fragments, one of mass x and the other of mass y. The breakup rates in eqs. (10.3)
are related by a(x) = f£ F(y,x — y) dy, and for the case of binary breakup, f{x\y) =
f(y-x\y)·
The cluster size distribution that results from eq. (10.3) is determined by
the details of the kernels a(x) and f{x\y) (or, equivalently, by F(x,y)). With their
form left unspecified, it is only possible to find the general conditions for which
the fragment size distribution approaches a limiting form at long times. Conse-
quently, there has been considerable work on solving the rate equations for specific,
physically-motivated breakup kernels. Exact solutions have been found for models
which are meant to account for situations such as ore comminution, powder crush-
ing, and fly ash formation. In particular, Ziff and coworkers have recently solved the
rate equations of fragmentation for a relatively wide class of modelsJ 58_6 °] Their
work is notable for the general insights gained which parallel recent progress made
in the investigation of aggregation phenomena. An illustrative example of this type
of solution will be given in the next subsection.
By focusing on specific examples, however, it becomes difficult to appreci-
ate the range of possible behaviours. I will therefore outline a scaling approach to
fragmentation phenomena, in which universal features of the fragment size distri-
bution emerge clearlyJ63,64^ In order to apply scaling, I will henceforth specialize
to the case of systems with breakup rates that are homogeneous functions of the
fragment size. Homogeneity implies that the relative size, rather than the abso-
lute size is the basic parameter which characterizes a breakup event. Therefore the
overall breakup rate depends on the mass of a fragment as a(x) = χ λ , where λ is
termed the homogeneity index. According to the discussion of the last section, we
expect that λ is positive in general. Homogeneity also implies that f(x\y) depends
only on the ratio of the mass of the product to the mass of the initial cluster, i.e.,
f(x\y) oc y~l b(x/y). From the experimental data shown previously, b(x) is often
described by a power-law form in the small-x limit. By definition, the integral
2. Continuous fragmentation 329

/0°° b(x) dx equals the average number of fragments produced in a single breakup
event, and mass conservation imposes the condition /Q1 X b(x) dx = 1.
Ostensibly, the restriction to homogeneity is not very restrictive, as a wide
range of physical systems appear to be described by homogeneous kernels. Possible
limitations associated with the restriction to homogeneity which will be discussed
in section 10.4. In the discussion that follows, we will outline both exact and scaling
solutions to fragmenting systems with homogeneous breakup rates.
10.2.2 Illustrative exact solutions
Since the rate equations for continuous fragmentation are linear, they are
soluble, in principle, for arbitrary breakup kernels. Ziff and McGrady have outlined
a formal procedure which yields the general solution to the rate equations for the
case of binary breakupJ 58 ' 59 ! A salient feature of these solutions is that for fixed x,
the concentration c(x,t) decays purely exponentially in time as t —> oo. This fact
can be seen directly from the rate equations. At sufficiently long times, the gain
term in the rate equations can be neglected for any finite value of x. Consequently,
the leading behaviour for c(x,t —> oo) is proportional to e~ e ^*. The role of the
gain term is to introduce power-law modifications to this asymptotic exponential
behaviour. Building on this insight, one way to obtain the solution for all times is
to write the full solution as a product of the long-time solution times a power series
in £, in which each coefficient is a function of x. This leads to recursion relations
for the series coefficients which can be solved in closed form for various cases.
To illustrate the important features of the fragment size distribution, let us
consider one very simple class of models in which there is binary breakup, with the
breakup rate depending only on the size of the object breaking up, and not on the
sizes of the product fragments.^ Using the description given in eq. (10.3b), this
corresponds to F(x, y) depending only on the combination x+y. For a homogeneous
system, one therefore has F(x,y) oc (x + ?/) λ_1 . Models of this kind have been
invoked to describe the breakup of polymer chains under the action of shear or
by chemical degradation, for the case where λ = 1. This describes the situation
where each bond in the chain breaks at a constant rate. For general values of λ, the
solutions to the rate equations illustrate how the overall breakup rate manifests
itself in the form of the cluster size distribution.
For F(x,y) = (x 4- y)x~x, eq. (10.3b) becomes

^ψΐ = -χ" c(x, t) + 2 j y^c(y, t) dy, (10.4)


X

or by differentiating with respect to x,


cxt{x, t) = -xxcx(x, t) - (A + 2)xx~lc(x, f), (10.5)
where the subscripts denote partial differentiation. In view of the limiting asymp-
totic exponential form for c(x,t), we attempt a solution of the form
c{x,t) = A(t)e-mx\ (10.6)
330 S. Redner / Fragmentation

c(x,t) c(x,t)

I i 1— i 1 1 I i i i I
1 101 10 2 10 3 .2 .4 .6 .8
t x

Figure 10.3 Sketch of the fragment size distribution, c(x,i) = 2te~tx , based on the exact
solution to the rate equations for the case of binary breakup with F(x, y) = (x + y) (from
ref. [58]). Shown is, (a) the distribution as a function of time for various values of mass #, and
(b) the distribution as a function of x at various times.

This yields soluble ordinary differential equations for A(t) and #(£), from which
one obtains a special solution for c(#,£),

c(x, t) = (l + * / * ) 2 / V ( ' + ' ) x \ (10.7)

where s = B(0). By the linearity of the rate equations, the general solution for
c(x,t) can be written as a linear combination, that is, an integral transform of
eq. (10.7), which, in turn, can be expressed in terms of special functions.
To gain some intuition for the fragment size distribution, consider c(x,t)
in the long time limit. This gives the asymptotic behaviour,

c(x, t) oc t2/x exp{-txx). (10.8)

A sketch of this fragment size distribution is shown in fig. 10.3, both as a function
of time for fixed size, and as a function of size at fixed times. Qualitatively, the
behaviour is in accord with simple-minded intuition. The population of a given
(small) size, x, initially grows due to the breakup of larger clusters. Eventually,
however, this size population decays when the production of x diminishes due to
the depletion of larger clusters. This decay is asymptotically exponential, with the
decay time varying inversely in the particle size. For the distribution at fixed time,
there is a steepening near the origin as a function of time, reflecting the eventual
predominance by very small clusters. This small mass tail often has a power-law
2. Continuous fragmentation 331

form, as discussed previously. Notice also that this asymptotic form for c(x,t)
becomes pathological as λ —> 0. This is a signal of the shattering transition, which
will be discussed in a later section.

10.2.3 Scaling formulation


As discussed in the introduction, the scaling approach is a simple but very
powerful conceptual tool to analyze the solutions to the rate equations. According
to scaling, the ratio of the size to the typical size is the only parameter which
characterizes a fragmenting system in the scaling limit. This scaling behaviour is
already evident from the exact solutions treated above. There are a number of
important reasons to emphasize scaling solutions. First, scaling generally provides
the simplest route to the asymptotic solution of the rate equations, especially in
complicated situations, where exact solutions require considerable technical exper-
tise. This simplification follows because the scaling ansatz reduces a two-variable
problem to a single variable problem. This reduction is an important motivating
factor for using the scaling approach for dynamic phenomena in general. Second,
a scaling solution is universal in that it is independent of initial conditions. Thus
scaling provides a universal classification of the solutions to the rate equations for
an encompassingly-wide wide class of fragmenting systems.
The scaling ansatz for the cluster size distribution can be written as^67,73~76^

c{x,t) = s'2<l)(x/s{t)), (10.9)

where s(t) is the typical (time-dependent) cluster mass, and the exponent —2
is required by mass conservation. Thus the cluster concentration at long times
depends on the ratio of the mass to the typical mass, rather than being a function
of mass and time separately. A scaling ansatz of this general spirit forms the basis
for the asymptotic analysis for a wide variety of dynamic phenomena, and has had
extensive success in processes such as aggregation.
In our scaling approach, it is easiest to first compute the moments of the
fragment size distribution, and then reconstruct the distribution, rather than com-
puting the distribution directlyJ 63,6 ^ For this purpose, we define the a t h moment
of the cluster size distribution, MQ(£), and the a t h moment of the scaling function,
raQ, by
OO OO

Q
Ma(t) = I x c{x, t) dx, ma= f xa φ{χ) dx. (10.10)
0 0

These are just the Mellin transforms^ of c(x, t) and φ(χ), respectively (except for
a trivial shift of unity in the definition of a ) . From these definitions, M0(t) is the
total number of clusters, and M\{t) is the total mass in the system at time t. In
these definitions, there are two free parameters which arise from the amplitudes of
s(t) and φ in the scaling ansatz. Without loss of generality, we choose m0 = πΐχ = 1
in order to fix these two free parameters.
332 S. Redner / Fragmentation

For systems which obey scaling, the "bare" and "scaled" moments are
related by
Ma(t) = maa{t)a'1. (10.11)

Thus all the moments of the cluster size distribution are accounted for by a single
size scale s(t), and the special case a = 0 yields,

s(t) = m0/M0(t). (10.12)

Thus for a system with a unique typical size scale, the average size is inversely
proportional to the average number of clusters. The relation between the scaled
moments ma and the bare moments Ma(t) follows solely from the ansatz (10.9),
therefore this relation is common to fragmentation, aggregation, and other dynamic
phenomena which are described by scaling.
A crucial step in our scaling analysis is to relate the moments of the cluster
size distribution to the corresponding distribution itself. This relation is embodied
by the following correspondence. First, by scaling, we compute the moments ma
directly for a discrete set of equidistant a values. We then invoke "smoothness",
in which we assume that the form of the moments defined on the discrete set {a}
continues to hold for all real values of a. Finally, the functional form of the scaling
function for the cluster size distribution, φ(χ), is determined by computing the
inverse Mellin transform of ma. As an example, if the moments ma are finite for a <
ac and infinite for a > a c , then it is interpreted that φ(χ) asymptotically behaves
as the power law, φ(χ) oc x~l~a*. Although this procedure is not mathematically
precise, the correspondence between ma and φ(χ) is physically plausible, and it is
also supported by available exact solutions. This general procedure underlies our
ensuing discussion for the asymptotic form of the fragment size distribution.

10.2.4 Scaling solutions to the rate equations


To determine the asymptotic solution to the rate equations, we now substi-
tute the scaling ansatz into eq. (10.9). This separates the mass and time dependence
of the rate equation to yield two separate scaling equations for the time dependence
of the typical fragment size, and the functional dependence of the fragment size
distribution on scaled mass £ = #/sJ 63,64 J

ω{2φ(ξ) + ξφ'(0] = -ξΧΦ(0 + /φ(ν)νλ~^β]άη, (10.13)

i(i)s(<)" ( 1 + A ) = -ω. (10.14)

Here ω > 0 is the separation constant whose value depends on the normalization
of mo and mi, and the overdot denotes the time derivative. Prom eq. (10.14), the
2. Continuous fragmentation 333

typical cluster size has the time dependence,

t~1/x, for λ > 0 and t -► oo,

s(t) e _ü,i , for λ = 0 and t -* oo, (10.15)

(*c - *) 1 / | λ | , for λ < 0 and t < tc.

For a smaller value of λ the breakup rate of small clusters plays an increasing role.
Therefore the typical size decreases more rapidly for a model with a smaller value of
A. When λ becomes less than zero, the fragmentation process is dominated by the
rapid breaking up of the very smallest clusters. This phenomenon is manifested by
the fact that the scaling approach predicts that the typical size vanishes in a finite
time. This is the signal of the shattering transition, whose existence and properties
will be treated in the next subsection. For the present discussion, we consider the
case A > 0, where scaling can be shown to hold.
To find the asymptotic solution to eq. (10.13), we first convert it to a rela-
tion involving the moments of (/>(£), by multiplying both sides by £ a and integrating
over all £. After a number of straightforward steps, one obtains a linear recursion
relation for the moments of the scaling function,
1- a
rna+x = ω - raa, (10.16)
JL/a — 1

where La is the a t h moment of the reduced breakup kernel


1
La = f xab(x)dx. (10.17)
o
The explicit dependence on the kernel is only contained in La. Consequently, the
results which follow from this recursion relation will be universal to the extent that
they depend only on the homogeneity index that appears in the overall breakup
rate and on the limiting behaviour of L Q .
To obtain the time dependence of the moments, we start with the original
rate equation and convert it to a moment relation by following exactly the same
steps that led to eq. (10.16). This yields the "bare" moment relation,

Ma = (LQ - 1) Ma+X. (10.18)

Starting with Mx(t) = 1, eq. (10.18) gives Μχ_λ(*) = (Li_ A - 1)* + const., where
Li_A — 1 is a positive constant. Iterating this process leads to the asymptotic
solution,

M1.kX ~ I I ( L W A - l)fj oc tk, (10.19)


334 S. Redner / Fragmentation

for the discrete set of equidistant index values 1 — kX. On the other hand, through
eq. (10.11) the scaling ansatz gives mi-k\ <* s~kX, which is also proportional to th.
Thus the correct temporal behaviour of the moments is reproduced directly from
the scaling ansatz. It is also worth emphasizing that from the theoretical view-
point, these negative moments of the cluster size distribution are the fundamental
dynamical quantities that characterize the fragmentation process. They play a role
analogous to that of the positive moments of the cluster size distribution in growth
phenomena such as aggregation.
From eq. (10.16), we now determine the asymptotic behaviour of ma and
then use the properties of the inverse Mellin transform^ to reconstruct the func-
tional form of the scaling function. These details are presented separately for
a —y ±oo, which from the properties of the Mellin transform, correspond to large
and small #, respectively.
10.2.5 Large-x limit
To obtain the moments for large a, choose a = &λ, with k a positive integer,
iterate eq. (10.16), and use ra0 = 1. This yields

™*A = Ü;*" 1 Π ^ 1
· (10.20)
l L
n=l ~ n\

For large fc, the product is dominated by the large-n factors. In this limit, the
form of Ln\ is determined by the behaviour of the breakup kernel b(x) for x —> 1,
i.e., in the limit of abrasion. For this limit, consider kernels of the general form
b(x) — 6(1) + 0 ( ( 1 — #) μ ), for x —> 1, where 6(1) > 0 and μ > 0 are constants.
This means that there is a smoothly varying probability for the production of
fragments whose mass is very close to the initial mass, with a limiting probability
6(1) corresponding to the case of no breaking. This form for the breaking rate gives

La = b(l)/a + 0 ( α " ( " + 1 ) ) , (10.21)

for large a. Using this result in eq. (10.20), one finds after several straightforward
steps,
α / λ
ma oc ( ^ ) aOW-W*- 1 ' 2 , for a -> oo. (10.22)
The non-trivial dependence on the breakup kernel appears only through
λ in the controlling factor, α α / λ , and through 6(1) and λ in the correction term,
α (6(ΐ)-ΐ)/λ fpjug universality of ma leads, through the inverse Mellin transform, to
the following universal expression for the cluster size distribution

φ(χ) ~ xb{1)~2 exp(-axx) x -> oo, (10.23a)

where a = 1/(λω). By comparing with the scaling ansatz, this gives

c(x,t) oc exp(— const.tx ) , (10.23b)


2. Continuous fragmentation 335

which coincides with the asymptotic form of the exact results obtained for binary
fragmentation models. The basic conclusion, therefore, is that scaling provides a
simple method for obtaining the universal asymptotic form for the cluster size
distribution in the large size limit, which agrees with the available exact solutions.
10.2.6 Small-z limit
In the small-mass limit, there is a lesser degree of universality, since the
small mass tail is not directly influenced by particles of the typical size. However,
there are now just two generic forms for φ(χ), whose applicability depends on
whether or not the moments L_ a diverge as a —► oo. Physically, L_ a diverging
corresponds to the production of a considerable fraction of arbitrarily fine dust in
a single breakup event, as might occur in an explosive process.
To obtain φ(χ) in the small-x limit, now requires the behaviour of ra_Q as
a —> oo. Accordingly, we choose a = 1 — kX in eq. (10.16) and iterate to arrive at
the counterpart of eq. (10.20), namely,

i
rai_*A = ω Λ . (10.24)
Π„ = ι ηλ
In analogy with the case of large positive index values, the k dependence of mi-k\
for large k is now determined by the limiting form of b(x) for x near 0, and there
appear to be two generic cases for this limiting form. One case is the general
situation of kernels which are cut off at small fragment sizes. This corresponds
roughly to the case of cleavage, as defined in the previous section. Such a process
may be modelled by a kernel with a finite lower limit, x0, for the minimum reduction
factor in a single breakup event. That is, b(x) = 0 for x < # 0 , with 0 < XQ < 1. From
eq. (10.17), Zq_a has the leading behaviour χ^α/α1+μ for large a. Substituting this
form into eq. (10.24) yields

x a x
rai-fcA ~ fcjf^fc o o · (10.25)

Thus the controlling factor of ma in the large a limit is

-ln#o o
m_ a ~ exp a a -> oo, (10.26)
2X

whose inverse Mellin transform yields the classical log-normal form for the control-
ling factor of φ(χ).

λ
φ(χ) ~ exp (In 2 *) (x -> 0). (10.27)
21n#o
In fact, this expression represents a strict lower bound for the small-mass limit of
the cluster-size distribution.
336 S. Redner / Fragmentation

The log-normal form can also be obtained from a simple multiplicative


argument which appears to capture some of the essence of repeated fragmentation
with a small size cutoffJ81-85^ For such a process, the mass of a given fragment
schematically evolves as x0 —> X\ —> x<i —> · · · —► XN, where the successive reduction
factor, rk = xk/xk-i, is a random variable with a well-behaved distribution. By the
central limit theorem, log xN = Σ/feLo1°8 rk will be normally distributed, so that xN
will be distributed log-normally. While this naive argument is appealing, it contains
the tacit assumption that fragmentation actually proceeds in discrete stages so that
each cluster has undergone approximately the same number of breakup events.
The essential difference, if any, between the cluster size distribution that arises
in this discrete dynamics and from a purely continuous process has not yet been
elucidated.
Nevertheless, the log-normal form does appear to describe quite adequately
the size distribution in a wide variety of geological situations, such as the distribu-
tion of rock sizes in boulder fields and the distribution of particle sizes in soils J1-4^
Typically, it is found that the log-normal provides a helpful representation of the
data whenever the size range of the population varies over several orders of mag-
nitude. For these situations, the log-normal is invoked as a truly fundamental
distribution, even though there may be only fair agreement between a given data
set and the log-normal form.
A second general class of fragment distributions arises for kernels in which
infinitesimal size pieces are produced in a single fragmentation event, as in de-
structive breaking. Although it is difficult to envision such a process occurring
repeatedly, the solution for this class of kernels completes the picture of the pos-
sible forms for the small mass tail of the fragment size distribution. The absence
of a cutoff in the breakup kernel is typified by the power-law form b(x) ~ xu for
small x. Thus from eq. (10.16), raQ diverges whenever La diverges. This divergence
occurs for a less than a critical value a c , which is less than 0, since ra0 is finite.
For a close to a c we keep only the leading term in eq. (10.16) to give

ma ~ La ™Qc+A a La. (10.28)


ω(1 -ac)
Since ma is proportional to L a , it follows that φ(χ) coincides with b(x). That is
φ(χ) has the power-law form
φ(χ) ~ x", as x -> 0. (10.29)
Such a power-law behaviour is seen in a variety of fragmentation experiments J52»55»62!

Thus for a kernel with no small size cutoff, the limiting form of φ(χ) now
decays as exp(—i/lnx), which should be compared to the more rapidly varying log-
normal bound, exp[—c(lnx)2] for a kernel with a small size cutoff. Eqs. (10.23),
(10.27) and (10.29) provide the asymptotic behaviours of φ(χ) for an encompass-
ingly wide class of breakup kernels. These differing forms for the fragment size
2. Continuous fragmentation 337

increasing
"flakiness"
fC(X,t)

DUSTY

SHATTERED

♦c(x,t)

GRAVEL

Figure 10.4 Phase diagram for linear fragmentation in the plane defined by the homogeneity
index λ and a loosely-defined parameter, the "flakiness" of the relative breakup rate. Small
flakiness corresponds to a vanishingly small probability of small flakes being produced in a single
breakup event, i.e., a small-size cutoff in &(#), while large flakiness corresponds to the opposite
limit of a power-law tail in b(x). In the phase plane there is a "shattered" phase for λ < 0, while
for λ > 0 there is a "gravel" phase for small flakiness and a "dusty" phase for large flakiness. The
fragment size distributions corresponding to these latter two phases are sketched (from ref. [64]).
338 S. Redner / Fragmentation

distribution may be considered as corresponding to different "phases" of fragment


products, as illustrated by the "phase diagram" of fig. 10.4.

10.2.7 Breakdown of scaling and shattering transition for λ < 0


For λ < 0, conventional scaling solutions to the rate equations do not
exist J51'60'64] Accompanying this anomaly is the intriguing phenomenon of "shat-
tering", in which mass is "lost" to a phase of zero mass particles as the fragmenta-
tion progresses. Thus λ = 0 is the dividing point which separates models for which
conventional scaling exists from models which undergo shattering. Although shat-
tering may be a mathematical pathology, since small fragments are generally more
difficult to break than large clusters at fixed energy density, there are analogies
with the inverse and physical process of gelationJ 68-72 ! Furthermore, the shatter-
ing phenomenon reveals important and intriguing features about the structure of
the solutions to the rate equations, which can be usefully studied by exact solutions
and by scaling approaches.
The physical manifestation of shattering is mass loss in the system, Μχ < 0
(more specifically, mass loss from the population of finite size clusters). As first
shown generally by Filippov,^ and by McGrady and Zifi60! for specific fragmen-
tation models, this occurs when A < 0. We now give a simple argument to show
that λ < 0 is the necessary and sufficient condition for shattering in a homogeneous
fragmenting s y s t e m . ^ To locate the shattering transition, we write eq. (10.18) for
a = 1 + e,
M1 = (L1+€-l)Mx+1+€. (10.30)
As e —* 0 + , La approaches unity from below (mass conservation), so that Μχ can
be nonzero only if M\+i+€ diverges as e —> 0. Without loss of generality, suppose
that the largest initial cluster mass is unity, so that there will be no cluster with a
mass larger than unity for t > 0. Then Ma is non-decreasing as a decreases, at any
fixed time. This fact, together with Μχ < oo, implies that Ma can diverge only for
a < 1. Thus a necessary condition for shattering is λ < 0.
To show that λ < 0 is also a sufficient condition for shattering, let us
assume the converse and derive a contradiction. For λ < 0, eq. (10.18) gives

M1+lM = (L1+]xl-l)M1. (10.31)

Under the assumption of no shattering, Mx is fixed, so that the right-hand side is a


negative constant. This implies that Μι+μ|, and consequently c(x,t) would vanish
at a finite time. On the other hand, we know directly from the rate equations that
c(x, t) must decay exponentially, or slower, in time for any x. This contradiction
implies that a necessary and sufficient condition for shattering to occur is λ < 0.
Some of the peculiar features of the cluster size distribution can already be
seen from the exact solution of the rate equations for binary breakup in which the
homogeneity exponent equals zero. This corresponds to a system on the borderline
2. Continuous fragmentation 339

Figure 10.5 Sketch of the size dependence of the exact fragment size distribution at fixed times
for binary breakup when F(#, y) ~ (x + y)-1 (eq. 10.32). The singular behaviour in the small-size
limit should be compared with the regular small-size behaviour for fragmenting systems with
positive homogeneity exponent (fig. 10.3b).

between scaling and shattering. For F(x, y) oc (x + y) x, the exact solution for the
cluster size distribution in the limit λ —> 0 isJ58^

2te~ [2tln(l/x)]n
c(M) = e - ^ z - 0 +
I Σ
nf0n!(n + l)!'
(10.32)

and the corresponding moments of this cluster size distribution are,

Mn(t) = / n exp[(l - n)t/{l + n)]. (10.33)

Thus the total number of fragments, M0(£), grows exponentially in time, in con-
trast to the power-law growth for M0(t) in the non-shattering regime. A sketch
of c(x,t) versus x for several values of t is shown in fig. 10.5. In the small mass
limit, a singularity in c(x,t) develops which signals the explosive growth of very
small clusters. This singularity nevertheless makes a finite contribution to the total
number of clusters.
It is also possible to develop an alternative scaling formulation for c(x,t)
which accounts for basic features of the fragment size distribution in the shattering
340 S. Redner / Fragmentation

regime. To justify this new scaling ansatz, recall that the conventional scaling
solution for φ(χ) (eq. (10.23)) yields c(x,t) ~ e~x '. The coefficient of t in the
exponential is a decreasing function of x for λ > 0, so that smaller clusters decay
more slowly. On the other hand, for λ < 0, one can show directly from the rate
equations that c(x, t) ~ e~l~ * for the monodisperse initial condition c(x, t = 0) =
δ(χ — /), where / is the initial mass of the fragments. That is, the coefficient of t
in the exponential is fixed, or equivalently the typical mass is time-independent
This anomaly reflects the fact that once the first few small clusters are produced,
they react so quickly that the initial mass remains behind as a virtually non-
reactive residue, and this residue determines the typical size. This is similar to the
behaviour in the inverse process of gelation, where small clusters remain behind
as the non-reactive sol phase in the presence of a very few, highly-reactive, large
macromolecules.
These arguments, together with an examination of the conventional scaling
ansatz and exact solutions in the shattering regime,^ lead to the following form
for the long-time behaviour of c(x, t) when λ < 0,

c(x,t) = T(trM)(/){x/l), (10.34)

where T(u) is a scaling function. This resembles the conventional scaling ansatz,
with s(t) being replaced by / in the scaling function φ. Since we have shown that
c(#, t) has the time dependence e~z ', it now remains to determine the functional
form of φ(χ) to complete the description of the fragment size distribution. By a
careful examination of the moment equation for λ < 0, it is possible to show that
φ(χ) has the universal form χ'λ'~2. Thus we conclude that in the shattering regime,
c(x, t) has the universal behaviour

φ;,*)~β-Ηλ|<;τ|λ|-2. (10.35)

This functional form can also be justified by a very simple argument which
is based on the physical picture that most of the mass in the system remains with
the nearly-const ant initial mass residue. If the concentration of this residue was
strictly constant, then the dynamics of the system could be described by the rate
equation (10.3), augmented by the presence of a constant source of clusters of mass
unity. For this modified rate equation, the moment relation becomes^64!

ma_,A| * — i y - (a > 1 - |A|). (10.36)

This turns out to coincide with a detailed calculation for moments of the cluster size
distribution for irreversible fragmentation in the λ < 0 regime. Thus we conclude
that for λ > 0, clusters of all sizes participate substantially in fragmentation,
leading naturally to scaling behaviour. However for λ < 0, only the very smallest
clusters fragment at any appreciable rate. The singular aspects of the fragment size
2. Continuous fragmentation 341

distribution are influenced by this population of vanishingly small clusters, while


the scaling portion of the distribution is determined by the nearly non-reactive
residue of initial-mass clusters. The features of the cluster size distribution also
occur in the inverse process of gelationJ 68-72 !
10.2.8 Comparison with comminution experiments
As mentioned at the outset of this chapter, there are many examples of
continuous fragmentation. Indeed, the preparation of a dispersed product by the
grinding of a coarse feed is an underlying theme in most mining engineering and
powder technology applications.^ 11-1 ^ The grinding or comminution of minerals by
milling machinery, in particular, provides an excellent laboratory for the compar-
ison of theory and experiments. Experimental conditions can be easily controlled,
and one has a reasonable physical picture of the microscopic breakup processes
that are involved in grinding. The detailed comparison between theory and exper-
iments therefore seems to be best developed in the literature on comminution. In
the comminution of ores, the important practical question to address is the relation
between the energy input to the system and the degree of size reduction of the ore.
More precisely, for a given distribution of feed and for a fixed grinding time, the
issue of practical relevance is the size distribution of the ground material.
In comminution by ball or rod milling, for example, the initial feed is placed
in a cylindrical container, which is typically of the order 102 cubic meters in volume,
and spins at a rate that is typically of the order of 102 revolutions per minuteJ11^
The grinding action is accomplished by the presence of a very hard milling medium,
such as steel balls or steel rods. As the container spins, the collision of the ground
material with the steel is the primary mechanism for size reduction. The grinding
time is typically of the order of minutes to many hours, depending on the degree
of size reduction needed. A wide variety of geometries and milling processes have
been devised, depending on the specific application that is required. The large
degree of mixing inherent in a tumbling grinding mill suggests that material is
homogeneously distributed in the mill. (However, for steady-state operation, mills
have been designed to produce size segregation, with smaller fragments close to the
output endJ11!) If spatial homogeneity does apply, then one of the basic assump-
tions of the rate equations is satisfied. Thus for milling, at least, the mean-field
rate equations do provide an appropriate starting point.
On the basis of these types of comminution experiments, a number of basic
conclusions have been drawn. First, for processes such as ball or rod milling, the
fragmentation process is linear to a high degree of accuracy.^12'55! This linearity
has been tested by verifying that the concentration of material within a given size
range is depleted at an exponential rate. This can be accomplished by studying the
subpopulation of the largest initial clusters, or alternatively, following the evolu-
tion of a radioactively marked subpopulation of smaller clusters in the initial feed.
Experiments on many types of ground media also indicate that the overall breakup
rate of a cluster of size x, a(x), is a decreasing function of x, whose precise form
342 S. Redner / Fragmentation

depends on the material being ground and on the milling conditions. For one class
of examples where careful measurements have been performed, a(x) is found to
vary as # λ , with λ ~ 0.3 — 0.4 for materials such as coal, limestone, and cement
clinker in ball milling conditions J52'55'61^ In addition, these experiments also suggest
that the relative breakup rate, f(x\y), has a power-law behaviour, f(x\y) ~ [xjyY
for a wide range of conditions. As shown in section 10.3, the corresponding frag-
ment size distribution should therefore have power-law tail at the small-size limit.
Experimentally there does appear to be a relatively universal behaviour, with the
cumulative mass distribution of mass less than x varying as # 7 , with 7 in the range
0.6 to 0.9.
Thus in strongly mixed milling media, the connection between the theo-
retical treatment of the rate equations and measurements is quite reasonable. For
a wide class of systems, it is even possible to extract a value for the homogeneity
exponent of the overall breakup rate, and also the nature of the relative breakup
rate from the apparent power-law small size tail of fragment sizes.

10.3 Discussion and outlook


Fracture and fragmentation underlie a wide variety of dynamical processes.
In this chapter, I have attempted to outline the basic features of continuous frag-
mentation and the underlying, relevant aspects of instantaneous fracture. It has
been found that over a wide range of energy input to the system, it is generally
harder to break up small clusters than large clusters. This corresponds to an overall
breakup rate which is a decreasing function of the mass, in the rate equations for
continuous fragmentation. For the distribution of fragment sizes that are produced
in a single instantaneous fragmentation event, there is considerable data which in-
dicates that for many materials and experimental situations, the cumulative mass
fraction which is smaller than mass x varies as xa, with a generally close to unity.
This corresponds to the relative abundance of fragments of size x varying as # 2 _ a / / 3
(eq. (10.1)). This power law form holds over several decades in the best experi-
mental situations, but clearly, there must be a cutoff in which a distribution at the
very smallest scale. Such a power-law form translates to a relative breakup rate
that does not have a small-size cutoff in the rate equations, a situation in which a
small-size cutoff must eventually play a prominent role.
A general treatment of continuous fragmentation has been given within
the framework of the rate equations. This is an approach of a mean-field charac-
ter, as fluctuations in the spatial positions of the clusters and in cluster shape are
ignored. Both of these approximations are quite drastic, and very little is known
about how to account for these fluctuations in a realistic and tractable manner. One
does expect that the rate equations will provide the correct asymptotic behaviour
for systems of sufficiently high spatial dimensionality, and that the mean-field de-
scription will break down below a critical dimensionality dc. Presently, there is no
understanding of how to determine this upper critical dimension. In contrast, for
3. Discussion and outlook 343

the inverse process of irreversible aggregation, there are well-established methods


for determining dc1 and it is known that dc = 2 for systems where the cluster reac-
tivity is mass independent,^86'87^ and that dc may be infinite for models in which the
reactivity is a strongly increasing function of massJ88! In fragmentation, however,
there is no obvious universal transport or mixing mechanism, akin to Brownian
motion in aggregating systems, with which one can begin to estimate the role of
fluctuations within a typical correlation volume. Thus it is an open question of how
to assess the degree of validity of the rate equations for describing the kinetics of
continuously fragmenting systems.
Nevertheless, the rate equations do provide a simple and comprehensive
treatment for the kinetics of continuous fragmentation. Within this approach, the
microscopic details of breakup events are accounted for by the overall breakup rate,
which specifies the likelihood for a fragment of a given mass to break (either by an
external agent or by collisions), and the relative breakup rate, which gives the size
distribution of products from a single breakup event. These fragmentation kernels
can be determined empirically for a wide range of systems. For many functional
forms for the breakup rates, it is possible to solve the rate equations in closed
form. A more general, yet very simple approach is based on scaling, in which it
is assumed that the fragment size distribution depends on mass and time only
through the scaling combination of the ratio of the mass to the typical mass. For
a homogeneous fragmenting system, scaling solutions exist when the homogeneity
index, λ, of the overall breakup rate is positive. Within the scaling regime, the
value of λ is the crucial parameter which characterizes the large-mass behaviour of
the cluster size distribution. Under very general conditions, the large mass tail of
the distribution is found to decay as e~tx . In the small-mass limit, the cluster size
distribution is determined by whether or not the relative breakup rate is cut off in
the small-size limit. With a cutoff, the cluster size distribution has a log-normal
tail, while in the opposite case, there is a power-law tail of very small dust-like
particles.
When λ < 0, a shattering transition occurs in which mass is lost to a dust
phase consisting of an infinite number of zero mass particles. It is possible to obtain
asymptotic dynamical behaviour through a scaling ansatz if one takes the initial
mass as the characteristic mass scale of the system. Owing to the rapid breakup
of very small clusters, the initial mass decays away extremely slowly, so that this
component of the distribution acts as a steady source for the remainder of the
fragment size distribution. This intuitive picture accords with exact solutions of
fragmenting systems with λ < 0.
While the rate equations are useful for treating a wide range of continuous
fragmentation phenomena, there are interesting questions of potential experimen-
tal relevance that should be addressed. At a simple-minded level, the validity of
a purely linear theory can be questioned in situations as violent as rock crushing
and in many other fragmentation phenomena. In rock grinding experiments, devi-
ations from the predictions of linear rate equations are found to occur in various
344 S. Redner / Fragmentation

casesJ55^ Both non-linear and non-local effects can be expected to play some role
in high energy fragmentation processes. When breakup is driven by high pressures,
there can be a transfer of stress across many fragments. When fragments possess
considerable kinetic energy, non-linear collisional effects may play a substantial
roleJ38! Some of these features are embodied in autogeneous millingJ11^ where the
breakup of fragments is caused only by the material undergoing comminution. A
satisfactory theoretical description has not yet been developed for this intriguing
non-linear process. A natural starting point worth developing is to consider the rate
equations with either time-dependent or concentration-dependent breakup rates.
Autogeneous milling also shares some features with an idealized collision-induced
fragmentation model, in which fragmentation is driven only by binary collisions be-
tween fragmentsJ64'89] In the rate equations for this process, fragmentation events
are driven both by an intrinsic breakup rate and by the rate at which two fragments
meet. The latter is described by the time rate of change of the concentration c(x11)
being proportional to the square of the overall density^64! This collision-induced
model exhibits a rich phenomenology, whose solution may provide a starting point
for treating autogeneous fragmentation.
Another potential example of a system undergoing non-linear fragmenta-
tion is the asteroid belt, if collisions between asteroids are the primary mechanism
that lead to continued breakup events. In addition to the apparent non-linearity,
there is inhomogeneity and possible self-organization in the spatial and velocity
distributions of the asteroids. In view of these facts, one can reasonably doubt the
applicability of a rate equation approach to discuss the size distribution of aster-
oids. Nevertheless, an early rate equation approach^ predicted that the density of
fragments of mass x varies approximately as x~18. This was in good agreement
with the observational data of that time, but current data reveals that it is inad-
equate to view asteroids as a single homogeneous populationJ10^ It remains to be
established whether a rate equation approach can capture the essential features of
asteroid dynamics and fragmentation.
In a more practical vein, it is important to develop approximations which
can take inhomogeneities of various types into account. While the assumption of
perfect mixing is most likely appropriate for ball milling comminution, perfect
mixing is clearly inadequate in many geophysical situations, where fragments may
remain in fixed spatial positions throughout the breakup process. Attempts to ac-
count for this type of spatial inhomogeneity have been attempted, but primarily at
a qualitative level. For the phenomenon of the formation of fragments (gouge) in
an earthquake fault, a "constrained comminution" model has been introduced,^
in which it is postulated that in a system of fragments with fixed relative positions,
fragmentation is most likely to occur between neighbouring fragments of the same
size. This model is based on the intuition that neighbouring fragments of dissimilar
sizes will tend to "nest", leading to a screening of tensile forces. This model leads
also to the question of how the rate of fragmentation depends on cluster shape.
References 345

Whether this shape dependence can be accounted for by averaging over fragment
shapes has not been seriously considered. This potential shape dependence also
appears to be ignore in the available experimental fragmentation data. Also note-
worthy is a related constrained fragmentation model of Derrida and FlyvbjergJ 91 !
which has a very different physical application in mind. In their model, there is
binary fragmentation of the unit interval, in which only one of the two product
fragments is available for continued breakup. This model was introduced to de-
scribe the evolution of the phase space of a disordered system, such as a spin glass,
as the temperature is lowered. This model is particularly intriguing for the rich
singularity structure exhibited the fragment size distribution.
In conclusion, fracture and fragmentation encompasses a wide range of
dynamic phenomena. The theoretical treatment of linear fragmentation processes
is fairly well-developed. However, there are many situations where non-linearity,
and inhomogeneities of various types will play an important role in the breakup
process. The understanding of the dynamics for these types of systems should pose
rich areas for new research.
I wish to thank Zheming Cheng for a most enjoyable collaboration on
some of the work reported here. This work has been supported in part by grant
#DAAL03-89-K-0025 from the Army Research Office. This this financial support
is gratefully acknowledged.

References
1. E.C. Dapples, Basic Geology for Science and Engineering (Wiley, New York,
1959)
2. Models of Geologic Processes; An Introduction to Mathematical Geology, Ed.
P. Fenner (American Geological Institute, Washington, 1969)
3. P. Habib, Soil and Rock Mechanics (Cambridge University Press, Cambridge,
1983)
4. T. H. Wu, Soil Mechanics (Allyn and Bacon, Boston, 1966)
5. See, e.g. , Geology and Geophysics of the Moon, Ed. G. Fielder (Elsevier,
Amsterdam, 1972)
6. J. Guest, Geology of the Moon (Wykeham Publications, London, 1977)
7. W. K. Hartmann, Icarus 10, 210 (1969)
8. J. S. Dohnanyi, J. Geophys. Res. 74, 2531 (1969)
9. C. J. Cunningham, Introduction to Asteroids (Willmann-Bell, Inc., Richmond,
VA, 1988)
10. See, e.g. , the collection of articles in Asteroids Ed. T. Gehrels (University of
Arizona Press, Tucson, 1979)
11. E. G. Kelley and D. J. Spottiswood, Introduction to Mineral Processing (Wiley,
New York, 1982)
12. L. G. Austin, R. R. Klimpel, and P. T. Luckie, Process Engineering of Size
Reduction: Ball Milling (Society of Mining Engineers, New York, 1984)
346 S. Redner / Fragmentation

13. L. G. Austin and R. R. Klimpel, Ind. Eng. Chem. 56, 18 (1964)


14. E. J. Pryor, Mineral Processing (Elsevier, Amsterdam, 1965)
15. A. J. Lynch, Mineral Crushing and Grinding Circuits (Elsevier, Amsterdam,
1977)
16. H. Mark and R. Simha, Trans. Faraday Soc. 35, 611 (1940)
17. E. Montroll and R. Simha, J. Chem Phys. 8, 721 (1940)
18. P. J. Blatz and A. V. Tobolsky, J. Phys. C 49, 77 (1945)
19. R. Simha, L. A. Wall, and P. J. Blatz, J. Polymer. Sei. 5, 615 (1950)
20. H. H. Jellinek and G. White J. Polymer. Sei. 6, 745 (1951)
21. A. M. Basedow, K. H. Ebert, and H. J. Ederer, Macromolecules 11, 774 (1978)
22. R. M. Ziffand E. D. McGrady, Macromolecules 19, 2513 (1986)
23. C. A. Sundback, J. M. Beer, and A. F. Sarofim, Proc. 20i/l Int. Symp. on
Combustion (The Combustion Institute, Pittsburgh, 1985)
24. G. Daccord and R. Lenormand, Nature 325, 41 (1986)
25. D. Dunn-Rankin and A. R. Kerstein, Combust, and Flame 69, 193 (1987)
26. A. R. Kerstein and B. F. Edwards, Chem. Engr. Sei. 42, 1629 (1987)
27. M. Sahimi and T. T. Tsotsis, Phys. Rev. Lett. 59, 888 (1987)
28. R. Shinnar, J. Fluid Mech. 10, 259 (1961)
29. D. Ramkrishna, Chem. Engr. Sei. 29, 987 (1974)
30. L. A. Tavlarides and M. Stamatoudis, Adv. Chem. Eng. Sei. 11, 199 (1981)
31. P. Becher (ed.), Encyclopedia of Emulsion Technology, vol. 1 (Marcel Dekker,
New York, 1983)
32. A. P. Siebesma, R. R. Tremblay, A. Erzan, and L. Pietronero, Physica A 156,
613 (1989)
33. See, e.g., D. L. Turcotte, J. Geophys. Res. 91, 1921 (1986), for a recent review
34. B. R. Lawn and T. R. Wilshaw, Fracture of Brittle Solids (Cambridge Univer-
sity Press, Cambridge, 1975)
35. S. M. Weiderhorn, Ann. Rev. Mater. Sei 14, 373 (1984)
36. R. W. Davidge Mechanical Behavior of Ceramics, Cambridge Solid State Sci-
ences Series (Cambridge University Press, Cambridge, 1979)
37. See, e.g. the collection of articles in Fracture Vol. I - VII, H. Liebowitz ed.,
(Academic Press, New York, 1984)
38. See e.g., Fragmentation, Form and Flow in Fracture Media, R. Englman and Z.
Jaeger eds., Annals of the Israel Physical Society, vol. 8 (Adam Hilger, Bristol,
England, 1986)
39. P. Rosin and E. Rammler, J. Inst. Fuel 7, 29 (1933)
40. R. Schuhmann, Jr., Trans. AIME/SME 217, 22 (1960)
41. J. J. Gilvarry, J. Appl. Phys. 32, 391 (1961)
42. J. J. Gilvarry and B. H. Bergstrom, J. Appl. Phys. 32, 400 (1961)
43. J. J. Gilvarry and B. H. Bergstrom, Trans. AIME/SME 193, 484 (1961)
44. A. M. Gaudin and T. P. Meloy Trans. AIME/SME 223, 40 (1062); 223, 43
(1962)
45. R. R. Klimpel and L. G. Austin Trans. AIME/SME 232, 88 (1965)
References 347

46. A. Fujiwara, G. Kamimoto, and A. Tsukamoto, Icarus 3 1 , 277 (1977)


47. M. A. Lange, T. J. Ahrens, and M. B. Boslough, Icarus 58, 383 (1984)
48. C. J. Allegre, J. L. Le Mouel, and A. Provost, Nature 297, 47 (1982)
49. R. F. Smalley, Jr., D. L. Turcotte, and S. A. Solla, J. Geophys. Res. 90, 1894
(1985)
50. D. L. Turcotte, R. F. Smalley, Jr., and S. A. Solla, Nature 90, 1894 (1985)
51. A. F. Filippov, Theory Probab. and its Appl (Engl. Transl.) 4, 275 (1961)
52. D. D. Crabtree, R. S. Kinasevich, A. L. Mular, T. P. Meloy, and D. W. Fuer-
stenau, Trans. AIME/SME 229, 201 (1964)
53. L. G. Austin, Powder Tech. 5, 1 (1971)
54. V. K. Gupta and P. C. Kapur, Powder Tech. 12, 175 (1975)
55. L. Austin, K. Shoji, V. Bhatia, K. Savage, and R. Klimpel, Ind. Eng. Chem.
Process Des. Dev. 15, 187 (1976)
56. L. G. Austin and P. Bagga, Powder Tech. 28, 83 (1981)
57. L. G. Austin, P. Bagga and M. Celik, Powder Tech. 28, 235 (1981)
58. R. M. Ziff and E. D. McGrady, J. Phys. A 18, 3027 (1985)
59. R. M. Ziff and E. D. McGrady, Macromolecules 19, 2513 (1986)
60. E. D. McGrady and R. M. Ziff, Phys. Rev. Lett. 58, 892 (1987)
61. T. W. Peterson, M. V. Scotto, and A. F. Sarofim, Powder Tech. 45, 87 (1985)
62. T. W. Peterson, Aerosol Sei. Tech. 5, 93 (1986)
63. Z. Cheng and S. Redner, Phys. Rev. Lett. 60, 2450 (1988)
64. Z. Cheng and S. Redner, J. Phys. A (1990) to appear
65. B. F. Edwards, M. Cai, and H. Han, preprint
66. M. Cai, B. F. Edwards, and H. Han, preprint
67. S. K. Friedlander and C. S. Wang, J. Colloid Interface Sei. 22, 126 (1966)
68. R. M. Ziff, J. Stat. Phys. 23, 241 (1980)
69. M. H. Ernst, E. M. Hendriks, and R. M. Ziff, J. Phys. A 15, L743 (1982)
70. F. Leyvraz and H. R. Tschudi, J. Phys. A 14, 3389 (1981); J. Phys. A 15, 1951
(1982)
71. E. M. Hendriks, M. H. Ernst, and R. M. Ziff, J. Stat. Phys. 31, 519 (1983)
72. M. H. Ernst, R. M. Ziff, and E. M. Hendriks, J. Colloid Interface Sei. 97, 266
(1984)
73. T. Vicsek and F. Family, Phys. Rev. Lett. 52, 1669 (1984)
74. M. Kolb, Phys. Rev. Lett. 53, 1653 (1984)
75. P. G. D. van Dongen and M. H. Ernst, Phys. Rev. Lett. 54, 1396 (1985)
76. K. Kang, S. Redner, P. Meakin, and F. Leyvraz, Phys. Rev. A 33, 1171 (1986)
77. A. A. Griffith Phil. Trans. Roy. Soc, Ser. A 221, 163 (1920-1921)
78. H. Rumpf, Powder Tech. 7, 145 (1973)
79. Y. Oka and W. Majima, Can. Met. Q., 9, 429 (1970)
«80. See, e.g., J. Mathews and R. L. Walker, Mathematical Methods of Physics
(Addison-Wesley, Reading, Massachusetts, 1970)
81. A. N. Kolmogorov, Doklady Akad. Nauk. SSSR 31, 99 (1941)
82. P. R. Halmos, Ann. Math. Stat. 15, 182 (1944)
348 S. Redner / Fragmentation

83. B. Epstein, J. Franklin Inst. 244, 471 (1947)


84. See e.g., J. Aitcheson, The lognormal distribution (Cambridge University Press,
London, England, 1957), and references therein
85. See e.g., E. W. Montroll and M. F. Shlesinger, in Nonequilibrium Phenomena
II: From Stochastics to Hydrodynamics, eds. J. L. Lebowitz and E. W. Montroll
(North-Holland, Amsterdam, 1984)
86. S. Redner and K. Kang, Phys. Rev. A 30, 2833 (1984)
87. L. Peliti, J. Phys. A 19, L365 (1986)
88. P. G. D. van Dongen, Phys. Rev. Lett. 63, 1281 (1989)
89. See also R. C. Srivastava, J. Atmos. Sei. 39, 1317 (1982), for a particular limit
of nonlinear fragmentation
90. C. Sammis, G. King, and R. Biegel, PAGEOPH 125, 777 (1987)
91. B. Derrida and H. Flyvbjerg, J. Phys. A 20, 5273 (1987)
349

Subject index

abrasion, 324 cavitation, 28, 41


absolute reaction rate theory, 292 cavities, 61
activation energy, 292 Cay ley tree, 211
aggregate, 270 central force, 104, 148, 173, 197, 232, 279
aggregation, 323 central limit, 202
analog model, 313 ceramic, 24, 33, 189, 215
anisotropy, 274 cermet, 26, 33
annealed disorder, 175, 261, 276 chain slippage, 307
anodic, 10 chemical effect, 53
aspect ratio, 12, 46, 50, 77 clay, 2, 77
cleavage, 12, 59, 324
backbone, 134, 245
clusters, 119
beam model, 105, 146, 173, 232, 280
coarse-graining, 94
Besquin relation, 9
cohesion, 269
bilaplacian, 231
cohesive forces, 16, 163
Bingham fluid, 4
cold-drawing, 27
block relaxation, 295
comminution, 328, 341
blue brittleness, 29
composite, 26, 46
bond-bending, 145
compressibility modulus, 90
bone, 29
concrete, 52, 115
Born model, 145, 173, 280
conductivity, 116
branched polymer, 66
confining pressure, 21
branching, 9, 18, 268
conjugate gradient, 172, 240
breakdown, 189
connectivity, 174
breaking characteristic, 5, 47, 239
constant-gap scaling, 138, 252
breaking force, 6
brittle, 5, 60, 72, 306 contact, 88
brittle-ductile transition, 5, 74, 219 continuous fragmentation, 322
brittle-tough transition, 5, 74, 219 continuum percolation, 208
bubble model, 185, 208 copolymer, 66
buckling, 7 correlation length, 115, 123, 182
Bui integral, 162 corrosion crack, 213
Cosserat, 90, 93, 173, 232, 280
capillary number, 80 covalent bonding, 33
carbon fiber, 26 crack arrest, 16, 179, 213
cataclastic flow, 21 crack nucleation, 63, 302
catastrophic rupture, 6 crack size distribution, 238
350 Subject index

craze, 27, 73 electrodeposition, 264


creep, 41, 55, 64 electronic structure, 166
critical exponent, 122 embrittlement, 10
critical point, 121, 160 endurance limit, 9
critical scaling, 208 energy balance, 168, 238
crossover, 191 energy release rate, 17, 161, 212
cup-and-cone surface, 61 entanglements, 307
cutting bond, 135 epoxy resin, 68
cyclic fatigue, 9, 29 Eulerian coordinate, 88
extreme statistics, 199
damage, 101, 178 extreme-value statistics, 237
dangling ends, 135, 245
Darcy's law, 263 failure life, 9
Deborah number, 79 fatigue, 9, 63
delamination, 46 fiber, 46, 66, 215
dendrite, 264 fiber bundle, 48, 209
dendritic growth, 263 fiber waviness, 48
dense branching, 264 fiberglass, 46
density-density correlation, 272 finite element method, 109, 170
dielectric, 221, 238 finite-size scaling, 128
dielectric breakdown, 179, 221, 231, 248, fixed point, 180
261, 270 flexural modulus, 150
differential stress, 21 Flory approach, 238
diffusion-limited aggregation, see DLA fluctuation, 159, 164, 202
dilatancy, 4 Fourier acceleration, 172
dilation in variance, 180 fractal, 53, 126, 251
dilute, 176 fractal dimension, 221, 272, 281, 295
dilute networks, 195 fractal pattern, 80
dilute-crack approximation, 236 fractography, 60, 221
dilution, 115, 168, 189, 230 fracture stress, 168
dimples, 62 fragmentation, 9, 321
discretization, 103, 109 friction, 55, 209
dislocation, 41, 59, 168, 215 fuse, 197, 231
disorder, 115, 159, 164, 189, 229 fuse model, 278
distribution of stress, 249
DLA, 174, 177, 213, 261, 270 Galilean invariance, 88
dominant-crack theory, 190 Gaudin-Schuhmann distribution, 326
drying, 53, 279 gelation, 338
duality, 107, 133 gels, 77
ductile, 5, 61, 72, 306 glass, 19, 221
durability, 54 glass transition, 27, 70
glassy phase, 35
effective medium, 189, 211, 245 glassy polymer, 313
elasticity, 87, 145 gouge, 21, 344
elastomeric, 68 grain boundaries, 35, 63, 317
electrical conductivity, 92, 128 Griffith, 17, 161, 212, 238, 323
Subject index 351

grinding, 341 lattice animal, 124


Gumbel distribution, 190 lattice trapping, 168
leap-frog, 167
hackle, 20 Legendre transform, 140, 252
Hadamard instability, 7, 165 Lennard-Jones potential, 167
hash method, 167 linear stability, 264
Hausdorff-Besicovitch, 126 linearity, 88
Hele-Shaw cell, 80, 262 localization, 7, 55, 60
Hertz law, 4 localization limit er, 165
heterogeneity, 41 loop current, 131
hierarchical model, 181, 190, 207, 248
homogeneity index, 328 mean field, 116, 159, 322
homogenization, 115 measure, 251
homopolymer, 66 Mellin transform, 331
Hooke tensor, 91 melting temperature, 70
Hooke's law, 3, 90, 293 meshing, 110
hydraulic fracturing, 80, 174 metal, 28, 59
hydrogen bonds, 304 mica, 77
hyperscaling, 125 microcavities, 12
hysteresis, 55 microcrack, 12, 55, 115, 168
milling, 344
ice, 29 mirror zone, 19
impact, 215, 324 mist zone, 19
inclination angle, 22 molecular dynamics, 160, 166
inclusion, 52 molecular interaction, 163
interatomic potential, 167 molecular weight, 68, 309
intergranular fracture, 63 morphological transition, 264
internal stress, 53 morphology, 35
invariant, 89 moving-boundary problem, 263
ionic bonding, 33 mud cracking, 296
isotactic, 69 Mullins-Sekerka instability, 264
Ivantsov parabola, 266 multifractality, 133, 249, 274

Jacobi's method, 172 Navier-Stokes equation, 88, 262


Joffe effect, 25 necking, 7, 61, 72, 309
joints, 24 Newton's law, 166
noise reduction, 275, 283
kinetics of crack growth, 293 non-linearity, 88
Kirchhoff law, 104 non-local damage, 165
non-Newtonian fluid, 4, 77, 268
Lüders band, 4, 28
normality law, 98, 101
Lagrangian coordinate, 88
Lame coefficients, 89 open branching, 264
Lame equation, 269, 279 order parameter, 122, 217
laminate, 209 over relaxation, 295
Laplace equation, 263
Laplacian growth, 263 Pascal, 6
352 Subject index

percolation, 119, 245, 251 scaling function, 241


percolation threshold, 168, 230 screened percolation, 155, 245
perfect plasticity, 4, 96 self-affine, 272, 283
plastic potential, 97 self-averaging, 202
plastic strain, 4, 18, 96 self-similar, 160, 271
plastic zone, 37, 163 semibrittle, 25
plasticity, 96 shattering, 323
Poisson ratio, 3, 55, 91 shattering transition, 333, 338
polyethylene, 68, 304 shear band, 73, 198, 297
polymer, 26, 66, 329 shear modulus, 90
polymer degradation, 321 shear thinning, 77
polymer fiber, 302 shocks, 9
porcelain, 25 sintering, 34
pores, 12 size distribution, 322
porosity, 37, 53, 78, 80, 212, 219 size effect, 55, 164
post failure, 6 slip, 21, 28, 59
power-law distribution, 230 slip band, 63
predictor-corrector method, 167 slip-weakening, 21
process zone, 37, 163 slippage, 305
psi, 6 smectite, 77
stick-slip, 23
quantum fluctuations, 168 strain-hardening, 4, 60, 72
quenched disorder, 175, 229, 276 strain-softening, 165
strength fluctuation, 207
radius of gyration, 272 stress concentration, 13, 302
random dilution, see dilution stress corrosion, 10, 178, 283
random resistor network, 115, 203, 250 stress intensity factor, 13, 37, 162
random walk, 270 surface cracking model, 317
rate equation, 322 surface energy, 36,161, 168, 212, 321, 324
Rayleigh waves, 18 suspension, 78
reinforcement, 54, 209
reliability, 203, 223 tectonic fault, 21
renormalization, 160, 180, 190, 306 tensile curve, 47
representative volume element, 115 tensile strength, 304
residual bonding, 215 theoretical strength, 36
residual stress, 179, 213 thermal effect, 53
rheology, 87, 94 thermoplastic, 66
Rice integral, 162 thermosetting, 68
river patterns, 60 thin film deposit, 296
rocks, 20 threshold, 119
roughness, 221 tip-splitting, 80, 266, 271
rubber elasticity, 27, 308 torsion tensor, 93
toughening, 75, 215
Saffman-Taylor, 80, 264 toughness, 9, 13, 213
sand-box method, 284 Tresca, 74, 98
scaling, 121, 190, 233, 323, 331 twin, 28
Subject index 353

ultimate load, 47
ultimate stress, 26
uniform distribution, 230
universality, 123
unstable-crack approximation, 236

vacancies, 168
van der Waals, 33, 304, 307
Verlet method, 167
viscoelasticity, 5, 77, 82, 94, 269, 300
viscoplasticity, 5
viscosity, 70, 80
viscous fingering, 80, 261
von Mises, 74, 97, 233, 280

Wallner lines, 19
Weibull distribution, 10, 39,177, 190, 207,
231, 326
whiskers, 49
wood, 29
work-hardening, 99

yield point, 4, 72
yielding criterion, 96, 280
Young modulus, 3, 55, 74, 91

//-model, 271

Potrebbero piacerti anche