Sie sind auf Seite 1von 226

Partially Ordered Systems

Editorial Board:
E.Guyon Lui Lam Dominique Langevin H.E. Stanley
Ecole Nonnale San Jose State University Laboratoire de Physique ENS Boston University

Advisory Board:
J. Charvolin W. Helfrich Patrick A. Lee
Institut Laue-Langevin Institut fUr Theorie der Massachusetts Institute
Kondensierten Materie of Technology

John D. Litster David R. Nelson Martin Schadt


Massachusetts Institute Harvard University F. Hoffman - La Roche
of Technology &Co.

Springer Science+Business Media, LLC


Partially Ordered Systems
Editorial Board: L. Lam. D. Langevin

Solitons in Liquid Crystals


Lui Lam and Jacques Prost, Editors

Bond-Orientational Order in Condensed Matter Systems


Katherine 1. Strandburg, Editor

Diffraction Optics of Complex-Structured Periodic Media


VA. Belyakov

Fluctuational Effects in the Dynamics ofLiquid Crystals


E.!. Kats and V.V Lebedev

Nuclear Magnetic Resonance ofLiquid Crystals


Ronald Y. Dong

Electrooptic Effects in Liquid Crystal Materials


L.M. Blinov and VG. Chigrinov

Liquid Crystalline and Mesomorphic Polymers


Valery P. Shibaev and Lui Lam, Editors

Micelles, Microemulsions and Monolayers


W. Gelbart, A. Ben-Shaul, and D. Roux

Pattern Formation in Liquid Crystals


A. Buka and L. Kramer, Editors

Sands, Powders, and Grains: An Introduction to the Physics ofGranular Materials


Jacques Duran
Jacques Duran

Sands, Powders, and Grains


An Introduction to the Physics
of Granular Materials

Foreword by Pierre-Gilles de Gennes

Translated by Axel Reisinger

With 131 Illustrations

Springer
Jacques Duran Translated by
Directeur de Recherche Axel Reisinger
CNRS Lockheed Martin
University of Paris VI Lexington, MA
France USA

Editorial Board:
E. Guyon Lui Lam Dominique Langevin H.E. Stanley
Ecole Normale Department of Physics Laboratoire de Physique ENS Center for Polymer
45 RueD'Ulm San Jose State University 24 Rue Lhomond Studies
F-75005 Paris One Washington Square F-75231 Paris, Cedex 05 Physics Department
France San Jose, CA 95192 France Boston University
USA Boston. MA 02215
USA

Library of Congress Cataloging-in-Publication Data


Duran, Jacques.
Sands, powders, and grains: an introduction to the physics of granular materials / Jacques Duran.
p. cm.-(Partially ordered systems)
Includes bibliographical references and index.
ISBN 978-1-4612-6790-4 ISBN 978-1-4612-0499-2 (eBook)
DOI 10.1007/978-1-4612-0499-2
I. Granular materials. I. Title. II. Series.
TA418.78.D87 1999
620'.43-ddc21 98-31321

Printed on acid-free paper.

Translated from the French Introduction it la physique des materiaux granulaires, © 1997
by Editions Eyrolles, Paris, France.

© 2000 Springer Science+Business Media New York


Originally published by Springer-Verlag New York in 2000
Softcover reprint of the hardcover 1st edition 2000
All rights reserved. This work may not be translated or copied in whole or in part without the written
permission of the publisher (Springer Science+Business Media, LLC), except for brief excerpts in
connection with reviews or scholarly analysis. Use in connection with any form of information storage
and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now
known or hereafter developed is forbidden.
The use of general descriptive names, trade names, trademarks, etc., in this publication, even if the
former are not especially identified, is not to be taken as a sign that such names, as understood by the
Trade Marks and Merchandise Marks Act, may accordingly be used freely by anyone.

Production managed by Timothy Taylor, manufacturing supervised by Jeffrey Taub.


Photocomposed copy prepared by TechBooks, Fairfax, VA.

987 6 5 4 3 2 I

ISBN 978-1-4612-6790-4
Foreword

The physics of granular materials comes from an illustrious lineage. It includes


names like Coulomb during the reign of Louis XVI, Faraday and Reynolds in the
nineteenth century, and Bagnold, a remarkable Englishman who became enthralled
by the sands of the desert, perhaps even more so than T. E. Lawrence, to the point
that he resolved to understand its laws.
In spite of such great pioneers and their valiant efforts, the mechanics of powders
remains largely a mystery. Even seemingly trivial questions lack clear answers.
What is the weight distribution of a pile of sand dumped on the floor of an apartment
under constmction? Does it depend on the way it was dumped? No one really
knows.
Such questions happen to be quite important in a vaIiety of technical fields. They
have applications to situations ranging from agricultural facilities storing corn in
silos to the space exploration of Saturn's rings-this "fascinating merry-go-round
of rocks," to quote Guyon and Ie Troadec.
I have just mentioned two contemporary authors. They recently wrote a book
entitled Le sac de billes (The bag of marbles), published in 1994 by Odile Jacob.
It is a fascinating introduction to these problems, as well to somewhat unusual
objects like porous materials. Yet, there is a need for a more advanced textbook
for the benefit of our science students. For the last few years, Jacques Duran has
taught a course in granular materials aimed at undergraduates in their senior year.
He exposed them to a few general principles-for instance, the existence of many
equilibrium states in the presence of friction. He was able to describe some key
factors with the help of simple experiments. Or perhaps I ought to say "almost
simple," for granular media can be deceptive. A naive theoretician like myself
may think that all the answers can be worked out with a couple of pounds of sand,
vi Foreword

a funnel, a few hoses, and a glass container. As it turns out, a thousand technical
details can upset an experiment and lead to nonsensical results. The physics of
granulars may not require costly hardware, but it demands enormous care. Nor is
it exempt of real dangers: Industrial reactors can generate violent explosions, even
when they are ostensibly empty.
I am thrilled to see French physics actively engaged in the field of grains and
powders, where almost everything is yet to be discovered. Within the last decade,
a number of highly creative teams have popped up in places like Paris, Lyon, and
Rennes. Jacques Duran leads one of those teams at the Pierre and Marie Curie
University in Paris. This book draws on his first-rate expertise. I personally had
the opportunity to read various drafts as it was taking shape. I learned a great deal
from them in the process. There was a time when we had to steer anyone interested
in sand to the classic text by Bagnold. From now on, Duran will be high on the
required reading list. This book will fill a very real need. I sincerely wish it great
success.
Pierre-Gilles de Gennes
Paris, France
Preface

"A body is liquid when it is divided into several


smaller parts that move separately, and it is solid
when all its parts are in contact."
Rene Descartes, Principles of Philosophy (1644-1647)
Granular materials were probably not on Descartes's mind when he wrote the sen-
tence quoted above, but in some sense they might as well have been. They are
indeed "divided into several smaller parts that move separately," although they
do not enjoy-far from it-the properties of a conventional liquid. And "all its
parts are in contact," although that does not confer on them the type of solidity
envisioned by Descartes. This ambiguity illustrates well the lack of understand-
ing, which has persisted until quite recently, of the properties of granular media,
in spite of their prevalence not only in nature but in many human endeavors as
well. Recent progress in several different areas of physics has fostered a critical
reassessment of the concepts passed on to us by the scientists and engineers of
generations past. As a result, a number of research groups dedicated to working on
granular materials have sprouted all over the world in the last decade. They have
produced a harvest of experimental observations, numerical simulations, novel
concepts, and theoretical models.
Faced with such an onslaught of new facts, a newcomer to the field is likely to feel
a bit overwhelmed. To make matters worse, this growing area of science has yet to
benefit from any standardization, even in matters of terminology. I myself came to
that realization while collecting material to teach an advanced course in the physics
of liquids. The purpose of this text is to provide a smoother introduction to this ex-
citing new area of research. Meant primarily as a teaching vehicle, it should answer
the needs of fourth-year college students, researchers just entering the field, and
young engineers interested in process engineering dealing with granular materials.
More generally, it is aimed at anyone with a sufficient theoretical basis who is inter-
ested in learning how to start from quite simple experiments and, from there, build
viii Preface

up concepts and theories about a medium whose behavior is most often surprising
and counterintuitive. My goal has been to gather, in as self-contained a manner
as possible and in a unified language, the background necessary to understand the
latest developments, as well as to present the rudiments of granular physics.
Given the current status of this young discipline, any progress is likely to come
from experiments. Henri Poincare once said: "It has been quite some time since
anyone has thought of getting ahead of experimentation or to construct the world
on the basis of a few hasty hypotheses." The remark is particularly relevant to the
physics of granular media. This book is rooted in that very principle. In this spirit, it
contains many descriptions of experimental devices and accounts of experimental
results obtained under conditions that are as controlled as possible. These provide
the backdrop for concepts, models, and ideas that are pushed to levels that some
might find too speculative. The adopted strategy, whose limitations must be borne
in mind, goes with the very nature of this discipline, which is still very much in its
infancy. In this context, this book discusses concepts and results as they are known
to us at the present moment. As such, it is only a snapshot, as our understanding
of this topic is bound to evolve, perhaps even undergo major revisions.
Lest it become a dry compilation of facts, writing a book of this kind entails
somewhat arbitrary choices that remain the sole responsibility of the author. On
more than one occasion, I felt compelled to leave out this particular result or that,
even though I may come to regret it in the future. I did so simply because it did
not readily fit in with the flow of the presentation or because it would have taken
me too far afield. I beg for those whose work is mentioned too cursorily or not at
all to forgive me. This by no means implies a negative judgment on my part. My
choice was guided exclusively by pedagogical concerns and my own concept of
the logical flow of the material. My decisions were often agonizingly difficult.
My thanks go to my colleagues and friends in my own laboratory (Eric Clement,
Jean Rajchenbach, and Touria Mazozi). I am also grateful to the members of the
Physics of Heterogeneous and Complex Matter Research Group affiliated with
the National Center of Scientific Research (CNRS) and with the European HCM
(Human Capital and Mobility) Network. I am enormously indebted to them not
only for much of the technical material in the book but also for infusing in it
their irrepressible enthusiasm which I hope will brighten the experience of the
reader. I acknowledge the careful reading of the french manuscript and comments
by Ko Okumura and VJ.-P. Faroux. Axel Reiseinger successfully carried out the
translation from the French edition. He kept within the spirit of the text while
suggesting numerous improvements and clarifications.
Finally, I owe a large debt of gratitude to Pierre-Gilles de Gennes and Etienne
Guyon, both of whom played an instrumental role in the early stages of the research
discussed here. I was fortunate to have their unrelenting support and encouragement
to write and publish the material in this book. It borrows a great deal from their
own work.
Jacques Duran
Paris, France
Contents

Foreword v

Preface vii

1. Introduction 1
1.1 Some Orders of Magnitude Defining the Problem 1
1.2 Economic Implications and Industrial Problems 3
1.2.1 Industrial Processing of Granulars . 4
Construction Materials . . . . . . . . . . 4
Processing Industries . . . . . . . . . . . 6
An Example: Casting by Sacrificial Polystyrene 7
The Agriculture Industry 10
1.2.2 Flow Problems . 10
1.2.3 Problems of Segregation . . 12
1.3 Granular Materials and Geophysics 14
1.4 A Brief Historical Review . . . . . 16
1.5 Prerequisites and Selected Bibliography . 18

2. Interactions in Granular Media 19


2.1 A Single Particle and Its Environment . 19
Laminar Drag . . . 20
Turbulent Drag .. 22
Granular Dendrites 23
x Contents

Humidity, Electrostatic Interactions, and Other Perturbations 24


Classification of Granular Materials and Definitions 25
2.2 Interactions between Two Particles. . . . . . . . . . . . 27
2.2.1 The Laws of Friction between Solids . . . . . . 27
The Three Fundamental Laws of Solid Friction. 28
A Microscopic Explanation 28
Gliding and Rotations: Frustrated Rotations 30
Rolling without Gliding 31
Gliding without Rolling 32
Transition from One Regime to the Other . 32
Stick-Slip Motion. . . . . . . . . . . . . . 32
2.2.2 Collisions and Deformations of Elastic Spheres 34
Frontal Elastic Collision . . . . . . . . . . . . . 34
Nonfrontal Elastic Collision and Rotation of Particles 36
A Ball Thrown Against a Wall . . . . . . . . . . . . 36
Nonfrontal Collision Between Two Elastic Spheres
with Friction . . . . . . . . . . . . . 39
The Tangential Restitution Coefficient 41
Penetration During Frontal Collision:
Hertz's Problem . . . . . . . . . . . 42
Inhomogeneous Spheres: The Soft Crust Model 44
2.3 A Single Particle on Top of a Granular Medium .. 46
2.4 Interactions Between Several Particles . . . . . . . . 49
2.4.1 The Laws of Friction in a Granular Medium 49
2.4.2 Bagnold's Number . . . . . . . . . . . 50

3. Fluidization, Decompaction, and Fragmentation 53


3.1 The Static Properties of a Granular Pile . . . 54
3.1.1 First Principle: The Role of Friction. . 54
The Stacking of Cannon Balls . . . . . 54
Indetermination of Solid Friction: Hysteresis . 56
Distribution of Stresses in a Granular Medium 59
Arch in Equilibrium Under its Own Weight .. 60
Arch Supporting an Evenly Distributed Load . 62
3.1.2 Stress-Strain Relations . 64
Identical Spherical Granules . . . . . . . . . 64
Granules of Different Sizes: Power Law and
Electrical Analogy . . . . . . . . . . . 64
3.1.3 Second Principle: Reynolds's Dilatancy . . . 65
Deformation of a Simple Parallelogram . . . 66
Deformation of a Row of Parallelograms Placed
Between Two Walls . . . . . . . . . . 69
3.1.4 Cylindrical Container: Janssen's Model. 70
Generic Model: The Silo Problem . 70
Specific Applications . . . . . . . . . . . 74
Contents xi

Cylindrical Container of Diameter D . 74


Two-Dimensional Container 75
3.2 Dynamic Properties of a Granular Pile. 75
3.2.1 A Column of Spheres Subjected
to a Vertical Vibration . . . . 76
Some Orders of Magnitude. . . 76
Mathematical Analysis of the Problem 77
Results: Fluidized Phase and Condensed Phase 79
Fluidization and Condensation as Functions
of Acceleration . . . . . . . . . . . . . . . 80
Fluidization and Condensation as Functions
of Height . . . . . . . . . . . . . . . . . 82
3.2.2 Two-Dimensional Stack of Frictionless Spheres . 84
Some Comments on Scaling . . . . . . . . . . . . 86
3.2.3 Two-Dimensional Stack of Spheres with Friction. 87
Generic Model 87
Levitation of Cylindrical Stacks . . . . . . . 89
Levitation of Two-Dimensional Stacks . . . 89
Experimental Observation of Decompaction
and Convection in a Two-Dimensional
Granular Structure . . . . . . . . . . . . . 90
Experimental Technique: Image Processing 91
Measurements of the Velocity
of Moving Particles. . . . . . . . . . . . . . . .. 91
Measurements of the Relative Motion
of Particles 91
Convection and Pile Formation 93
Threshold of Pile Formation and Decompaction 96
Dynamics of Pile Formation in Two Dimensions 96
Experimental Verifications .
of the Decompaction Model . . . . . . . 98
Short-Term Decompaction: Fragmentation 101
3.2.4 Fragmentation of a Stack in Guided Fall 104
Two-Dimensional Experiment 104
Theoretical Modeling . . . . . . . . . . 105
Fall Without Fragmentation . . . . . . 105
Where Do Fractures Initially Appear? . 107
Numerical Simulation of a Stack in Guided Fall 108
Distribution of Pressure in a Stack: Arch Effects 110
Self-Organization of Rotations . . . . . . . . . . 111
3.2.5 Surface Instabilities in an Extended Granular Medium. 111
Extended Three-Dimensional Stacks 114
Wavelength Dependence on Vibration Frequency 116
Extended Two-Dimensional Stacks 117
Summary 117
xii Contents

4. Granular Media in a State of Flow 119


4.1 A Sand Pile in Equilibrium: The Angle of Repose . . . . . . . . 119
The Embankment Angle of a Pile Made
of a Small Number of Particles . 123
Transition from Intermittent to Continuous Regime-
Power Laws . 123
Power Law for a Newtonian Fluid 126
Power Law for a Granular Surface 126
4.2 Avalanche Models. . . . . . . . . 127
4.2.1 Cellular Automaton Model (CAM) 128
The Principle . . . . . . . . . . . . 128
Implementations of the Cellular Automaton
Model (CAM) . 130
Lifetime and 1/f Noise . 131
The Statistics of Avalanches . 132
Piles Involving Large Numbers of Particles . 133
Piles Involving Small Numbers of Particles . 134
Relaxation of the Critical Angle-
Granular Temperature . . . . 136
4.2.2 Stick-Slip Model of Avalanches .. 138
Different Friction Models . . . . . 145
Burridge-Knopoff (BK) Pads. 145
Other Friction Laws F (v) . . . 146
A Few Remarks Concerning the Stick-Slip Model
of Avalanches. . . . . . . . . . . . . . . . . . 146
4.2.3 Avalanche Models Based on Coupled Variables 147
Upward Propagation of Perturbations 149
Simulation of Avalanches .. 150
De Gennes's Modified Model:
Rotating Drum Experiment . . . . . . . . . . . . . . . 151

5. Mixing and Segregation 154


5.1 Introduction.................... 154
5.1.1 Oyama's Cylindrical Drum . 155
5.1.2 Potential Energy of a Heterogeneous Pile 156
Superposition of Stacks-Two Compact Stacks 157
Where Are the Defects Concentrated? . 158
5.2 Segregation by Vibration . 160
5.2.1 Simulation of Segregation by Size. 160
Two-Dimensional Model . 162
Three-Dimensional Model . . . . . 164
5.2.2 Experiments on Segregation by Vibration. 164
Experiments on Continuous and Intermittent Ascent . 166
Convection or Arch Effect? . 167
Contents X111

Convection and Segregation


in Three Dimensions . . . . . . . . . . . . . . 168
Convection and Segregation in Two Dimensions 168
5.3 Segregation by Shearing 171
5.3.1 A Single Particle in a Uniform Medium . . . 171
5.3.2 Segregation of Two Populations of Particles
of Different Size . . . . . . . . . . . . 176
Segregation Speed and Particle Size . . . 178
Segregation Speed and Rotation Velocity 178
Fractal Growth of the Central Cluster . . 179
5.4 Segregation in Oyama's Three-Dimensional Drum 181
5.4.1 ExperimentalObservations. 181
5.4.2 Savage's Model . . . . . . . 182

6. Numerical Simulations 184


6.1 Introduction . . . . . . . . . . . . . . . . . . . . 184
6.1.1 The Challenges of Numerical Simulation 185
6.1.2 The Different Simulation Methods. . . . 185
Hard Spheres and Soft Spheres . . . . . 186
Duration of Collisions and Chronology Problems 186
6.1.3 The Transition from a Discrete
to a Continuous Description. 187
6.2 Simulations of Collisions . . . . . . . . 189
6.2.1 Introduction........... 189
6.2.2 One-Dimensional LRV Procedure 189
6.3 Molecular Dynamics (MD) Simulations. 190
6.3.1 Elastic and Friction Forces . . . 191
Linear and Nonlinear Equations 191
Mechanical Analogies . . . . . 193
6.3.2 MD Collision Model . . . . . . 194
Linear Model of a Binary Collision 194
Nonlinear Model of a Binary Collision 196
The Detachment Effect . . . . . . 197
6.4 Simulation of the Dynamics of Contacts. . . . 199
6.5 Monte Carlo (MC) Simulations 202
Monte Carlo Technique for Stacking and Relaxation 203
6.6 Sequential Model of a Pile . . . . . . . . . . . . . . 206

Bibliography 209

Index 213
1
Introduction

1.1. Some Orders of Magnitude Defining the Problem


The physics of granular materials deals primarily with macroscopic objects. The
term "macroscopic" means here that the objects making up such materials must
at the very least be visible to the naked eye. This is in contrast to mesoscopic or
microscopic media. As it turns out, the very concept of granular material imposes
some further restrictions, raising the low end of the range somewhat above the
limits of our visual acuity. It is important to realize from the outset that the physics
of such objects has in most cases nothing whatsoever to do with the notion of
temperature, at least not as we normally understand it. To appreciate this, we need
only calculate the kinetic energy E k of a small bead of silicate glass-a basic com-
ponent of river sand. As we will see shortly, the physics of dry granular materials of
the type we are interested in involves objects that are typically 100 microns ({Lm)
in size or larger. For typical translation velocities of the order 1 cm/s, we find
that E k = ~mv2 ~ 10- 12 Joule. If this kinetic energy were due entirely to thermal
agitation, it would correspond to a temperature of 1011 K!
Furthermore, the loss /':,.E p of potential energy experienced by such a particle,
as it drops by a height equal to its own diameter d, is given by /':,.E p = mgd
(this corresponds to a situation where particles slide down past each other while
remaining in contact). We find that /':,.E p and Ek are roughly equal.
The conclusion is that the collective behavior of such a granular medium simply
cannot be described in terms of traditional Brownian motion. Some other mecha-
nism responsible for generating fluctuations is required if we insist on defining a
temperature of some kind [1], [2]. We will encounter many such examples in the

J. Duran, Sands, Powders, and Grains


© Springer-Verlag New York, Inc. 2000
2 1. Introduction

course of this book. Incidentally, it is instructive to calculate what the diameter of


a particle of this same material should be if it is to be the seat of any significant
thermal agitation at ordinary temperature (in the sense of E k "'" kT). The result is
about 1 /Lm, or some two orders of magnitude smaller than the size of the smallest
particles of interest here. This one observation alone constitutes a serious obstacle
to modeling the properties of granular piles by means of conventional thermal or
hydrodynamic variables. Except in a few very special cases, 1 it is clear that the
behavior of granular materials will be determined primarily by the more or less
preferential and collective displacement of the particles involved.
The fact that Brownian motion is inconsequential explains why granular ma-
terials almost invariably resist any attempt at mixing. As we will see, shaking a
collection of grains typically results in segregation, in other words, in separation
according to size, precisely the opposite to what is being sought. By contrast, it
is well known that in liquids, Brownian motion is the key mechanism responsible
for intimate mixing of the various components. The phenomenon of granular seg-
regation can be viewed as further evidence of the lack of thermal agitation with a
mean free path characteristic of the average distance between particles.
Both nature and industry make heavy use of a variety of granular materials cov-
ering an extremely broad range of shapes, sizes, micromechanical, and chemical
properties. In Chapter 2, we will mention a standard classification scheme based
on size, but we will not concern ourselves with shapes, nor will we discuss the
distribution laws of dissimilar media and associated measurement or identification
techniques. Readers interested in such issues are referred elsewhere [5], [6]. What
we will do is try to highlight the fundamental principles governing the physics of
these materials, and to that end we will focus on simple and well-characterized
objects. For reasons that will become clear later on, we will restrict our attention
almost exclusively to the physical properties and behavior of so-called d,y granu-
lar materials. Granted, a number of concepts that will come up during the course of
this book can, with some precautions, apply to more complicated objects, such as
pastes, muds, and other concentrated mixtures, in which particles interact with an
ambient fluid. Be that as it may, we will largely confine our study to particles with
idealized shapes (generally spherical or cylindrical) and whose micromechanical
propeliies are well understood. In order to convince the reader that granulars
present a serious challenge both from an economic perspective and on a more
fundamental physics level, we shall briefly review a few practical applications of
these materials as well as some basic problems which we will examine in more
detail in later chapters. In a somewhat different vein-of considerable importance
to human affairs, we might add-granulars of various degrees of complexity are at
the heart of the geophysical sciences. In fairness, geophysicists have often beaten
"conventional" physicists in defining many of the concepts to be discussed in these
pages.

lOne such case is the effect of a uniform vertical air flow, which keeps particles in suspension and
imparts to them random movements mimicking a classical Brownian motion [3]. It is also possible to
model the system by way of a thermal agitation of sorts when the particles undergo repeated mutual
collisions, as in a rapid flow [4].
1.2 Economic Implications and Industrial Problems 3

The physical laws governing the behavior of granular media actually apply to
objects whose dimensions cover several orders of magnitude. From grains, a few
hundred microns each, to ice floes drifting across the polar seas (over distances of
1000 km),2 not forgetting Saturn's rings (made of icy particles about 1 cm wide
distributed in a band roughly 1 km thick), the science of granulars covers at least
twelve decades of sizes.
The fact that aggregates seem to obey universal laws applicable over such a wide
range of dimensions and characteristics is a strong incentive to pursue fundamen-
tal studies in that area. For instance, phenomena of segregation and intermittent
blockages are pervasive in numerous industrial processes involving granular ma-
terials. Accordingly, the remainder of this chapter will be devoted to a brief and
selective review of a few techniques and processes that in one way or another are
subject to phenomena such as convection, segregation, blockages by arching, all
of which are routinely encountered in industry. These effects will be examined in
greater depth in subsequent chapters.

1.2. Economic Implications and Industrial Problems


As we have already pointed out, granular materials occupy a prominent place in
our culture. The worldwide annual production of grains and aggregates of various
kinds is gigantic, reaching approximately ten billion metric tons. Coal accounts
for about 3.5 billion tons ofthat total, cements and ordinary construction materials
for one billion tons, to which we can add equal amounts of sand and gravel. These
constitute what is generally referred to as low-cost raw materials. The processing of
granular media and aggregates consumes roughly 10% of all the energy produced
on this planet. As it turns out, this class of materials ranks second, immediately
behind water, on the scale of priorities of human activity. As such, any advance in
understanding the physics of granulars is bound to have a major economic impact.
The industrial technology used in the treatment of granular materials involves a
number of processes. First comes the extraction of ores, sands, and gravel, which
often relies on dredging. Next comes crushing and grinding, followed by separa-
tion, all of which are commonly used with low-value-added materials. Since raw
materials often account for more than 85 % of the total cost, it is easy to understand
why so little effort has been expended to improve the basic technology, which
often dates back to the nineteenth century. 3 Not much has been optimized, despite
the fact that methods of transport (fluidized beds, conveyor belts), storage (silos),
and mixing (e.g., cement trucks) figure in all stages of the industrial processing
of aggregates. Problems have received primitive solutions at best. The more spe-
cialized and developed arena of high-value-added materials include the cosmetic

2 Studies of the movements of icebergs, notably in the vicinity of port facilities, has been the object of
several research contracts with the Canadian Navy. It turns out to be closely related to the physics of
granular materials.
3 A recent report by
a US government agency stresses the obsolescence of the techniques used to process
granular materials. Significantly, it is entitled "Granular Materials: A Legacy of Neglect."
4 1. Introduction

and pharmaceutical industries, specialized chemistry, and the food industry, which
demand increasingly sophisticated processing technologies.

1.2.1. Industrial Processing of Granulars


Construction Materials
The construction industry (housing, hydraulic concrete needs, public works pro-
jects, and so on) consumes aggregates at the rate of seven tons per capita per year,
which generated revenues of some fifteen billion francs (three billion dollars) in
1994. 4 Next to water, aggregates are the most heavily used material. Natural
aggregates are of alluvial, volcanic, or sedimentary origin. After being extracted
from quarries or underwater deposits, they go through a series of steps that require
the handling of huge quantities of material in granular form. Among these opera-
tions are short-distance conveying (often by forced flow through pipes), scalping
(i.e., the separation of nonvaluable fine residues, which removes at least part of the
problem for subsequent processing), crushing and grinding, followed by sifting,
washing, hydrocycloning, and desliming (particularly to improve the properties of
sand and gravel). The sequence generally ends with storage and ultimately trans-
portation to the point of use. Most of these operations suffer from a troublesome
list of disruptive phenomena such as flow obstructions, formation of plugs, and
segregation, all of which we will have an opportunity to analyze in detail later on.
Industrial requirements are so diverse, and the properties of the corresponding
materials so complex, that it takes no fewer than 15 parameters to fully specify
an aggregate used, for instance, in the preparation of hydraulic concretes. They
describe a menu of characteristics like thickness, cleanliness, degree of crushing,
cohesion, coefficient of flatness, and a host of others. s This proliferation of em-
pirical coefficients attests to the inherent complexity of the industrial world, while
physicists try to reduce them to no more than two or three fundamental microme-
chanical parameters, perhaps because that is the most they can handle. It would
seem that physics has a long way to go to catch up with the real world. Yet-and
this is a very encouraging sign-the troublesome phenomena that plague industrial
processes out in the field (blockages, arches, segregation) show up readily even in
our idealized theoretical models. 6 In this regard, it is instructive to at least briefly
discuss some problems associated with the concrete industry.

4The numbers quoted here come from data published by Lafarge-Coppee, a world-wide leader in the
manufacture of all types of construction materials.
5Interestingly, the official list of parameters characterizing industrial aggregates does not include even
one of the micromechanical parameters considered of fundamental importance by physicists. In-
dustrial engineers do not care about things like the coefficient of elastic restitution of the coefficient
of friction. These omissions are ample evidence that practical problems in the real world often de-
mand a more complex, albeit more empirical, set of parameters than seem necessary in the idealized
atmosphere of a laboratory.
6This touches on a fundamental aspect of the scientific method, which can at times seem too reductive.
Physicists have to choose objects of study that are simplified enough to be amenable to analysis, yet not
so simplified as to lose any relevance to the real process they are intended to simulate, which is invari-
ably far more complex. We will confront several such choices during the course of this book. Making
the COlTect decision is a difficult task which, for the most part, rests on no more than sound intuition.
1.2 Economic Implications and Industrial Problems 5

FIGURE 1. Around 200 B.C., Appolonius of Perga invented the arrangement shown here
to tile a surface as fully as possible with circles of various sizes. This object enjoys the
property of self-similarity.

Paradoxically, how aggregates contribute to the strength of concrete was redis-


covered only very recently. It has only been a few years since the resilience of
bituminous or high-performance concretes was proven to be due in large part to the
characteristics of the aggregates used in their manufacture. Everyone knows that
concrete is essentially a mixture of aggregates and cement, and that it undergoes a
slow chemical transformation (on a scale of tens of years). Since it is a composite
material, the key to success is to minimize the influence of the more fragile binder
material, and maximize that of the much stronger granular material in its midst.
The practical solution is embodied in hypercompact arrangements, which consti-
tute a natural answer to the problem posed many centuries ago by Appolonius of
Perga (ca. 262, ca. 190 B.C.) [7]. It is illustrated in Figure 1.7
This idea is probably what inspired engineers to add granulars of various sizes,
spanning several orders of magnitude, to the mixtures they prepare to make high-
strength concretes. The finest ingredients in these mixtures are of submicron size;
they are called silica fume. Their cost is extremely high, but they have made it
possible to produce concretes which such high strength that the construction of
skyscrapers a full kilometer high has become technologically feasible. 8 Indeed, the
city ofTokyo is planning to erect just such structures. Nevertheless, itis not difficult
to anticipate the enormous problems encountered when trying to mix such a gran-
ular medium compactly, given the unavoidable phenomenon of size segregation.
It is only fitting to acknowledge that, although they were not familiar with the
dauntingly complex physics involved, engineers and specialists in the fields of

7Very much to the point, there exists an Arab saying that can be loosely translated as: "You think your
basket is full when you have stuffed it with oranges. In truth, it is full of void, for you can still add
nuts, and even chickpeas after thal."
8The quality of a high-strength concrete is often expressed as the height of a column with a one-
squaremeter base that withstands collapse. This maximum height has increased by more than one
order of magnitude in the last few years.
6 1. Introduction

concretes and aggregates have proven remarkably adept at building on their rich
experience and mastering the difficult art of devising fluid binders in which each
grain can find its ideal spot. No one can deny that systematic experimentation has
yielded the right answers, even though manufacturing costs are so high that appli-
cations are for the time being restricted to the construction of highly specialized
buildings. It is reasonable to expect that the possible contributions of physics to a
better understanding of the properties of construction materials may substantially
lower the costs of producing, conveying, and mixing them. These processes are
intimately related to problems of flow and segregation, which are to be taken up
in Sections 1.2.2 and 1.2.3.
We will refrain from treating here the principles of soil mechanics, which fall
largely in the domain of geophysics and geology, or the kinds of problems civil
engineers have long been accustomed to dealing with. There exist in this area
numerous models painstakingly put together, and with a respectable track record.
For instance, most soil-mechanics engineers tend to view granular materials as
continuous objects that obey the usual laws of classical mechanics (including
friction, stress-strain relations, plastic deformation, and the like), such as they
apply to homogeneous solids. As we will see, there are cases when this approach
raises serious questions. From a practical point of view, however, it does offer
the advantage of permitting simple calculations to predict the behavior of various
structures made of granular materials, such as roadways, with generally adequate
safety margins.
There is, however, one related phenomenon that deserves to be singled out,
because it shows up routinely in everyday life. We are referring to segregation
by shearing, an effect that is characteristic of most granular materials. Figure 2
describes the phenomenon in question, which civil engineers are quite famil-
iar with. They have to contend with it as they plan and manage construction
projects.

Processing Industries
Generally speaking, modern manufacturing methods of high-technology materials
rely heavily on the processing of granular materials. We are not about to review
each of the many applications, transformations, and various processes involving
granulars at some stage of the technology. One book would not be enough to
cover them all. We will simply give a few examples selected for the purpose of
highlighting the many-faceted problems faced by industry, and their connections
with the fundamental approach advocated in this book.
We start by making a clear distinction between two separate branches of the
chemical industry. On the one hand are industries that typically deal with large
quantities of materials in granular or powder form. Their problems are generally
similar to the ones we have already mentioned, such as transport through pipes,
storage, or granular segregation. On the other hand are the chemical or pharma-
ceutical industries where materials are prepared in relatively small quantities, but
with high-value-added materials. In this case, the demands in terms of purity and
reproducibility are obviously far more stringent. Accordingly, their manufacturing
1.2 Economic Implications and Industrial Problems 7

FIGURE 2. Grading the embankment on the side of a road with a bulldozer causes segre-
gation by shearing. The concentration of large rocks is much greater on the suti'ace of the
embankment than in the interior.

technology tends to be much more sophisticated. In this context, we will shortly


discuss a technique designed to get around-to the greatest extent possible-the
phenomenon of granular segregation.

An Example: Casting by Sacrificial Polystyrene


The technique discussed here has been used recently by the French automobile
maker Peugeot to produce engine blocks which would have been extremely difficult
to manufacture any other way. A number of electromechanical firms and steel
foundries have studied and developed similar processes. In principle, the method,
illustrated in Figure 3, is quite simple and of potentially low cost. Unfortunately,
its practical implementation runs into difficult problems because of a fundamental
property of granular materials, which we will examine in detail in Chapter 3, known
as convection by vibration. The sequence of steps required in this technique is as
follows:

• A polystyrene model of the object to be duplicated in metal form is prepared.


• The model is placed inside a container that is filled with fine sand.
• The whole system is shaken vigorously by means of a vibrating support placed
under the container.
• At the conclusion of this step, the sand occupies every bit of space between
the model and the container's wall. Highly compacted, it is now ready to act
as the mold.
8 1. Introduction

Polystyrene model
Vibration

FIGURE 3. Principle of casting by sacrificial polystyrene. The right-hand drawing shows


an industrial application of the principle. Several molds are placed in a container filled with
compacted sand. Granular convection phenomena in the sand can cause the structure to
deform prior to casting, resulting in a large number of rejects.

• Molten metal is poured through an opening connected to the polystyrene


model. The polystyrene instantly vaporizes away, leaving in its place the
liquid metal which solidifies into the appropriate shape.

In principle, the technique ought to produce nearly perfect results. One of its
advantages is that it eliminates the burrs formed at the joints between separate sec-
tions of conventional molds. Unfortunately, during the shaking process, granular
convection gives rise to a powerful central flux of sand which has a disturbing
tendency to deform-and sometimes to destroy outright-the polystyrene model.
This is a very serious problem that affects about 90% of the products, causing a
flood of rejects.
This problem is quite typical in the physics of granular materials. The solution
is relatively straightforward: Convection can be largely eliminated by using nearly
frictionless containers and granulars. But this ignores the fact that convection also
plays a useful role, as it promotes the filling of every nook and cranny around
the model. The trick is then to preserve the useful part of the convection while
minimizing or inhibiting the undesirable more violent component. Solutions do
exist, but they involve processes and materials so costly as to preclude their use
for mass production. A more pragmatic approach consists in finding the best
compromise between the dictates of physics and manufacturing costs.
This example is a classic illustration of the delicate balance that must be struck
between fundamental advances in physics and industrial requirements. It is often
not enough to propose clever solutions. We also have to contend with cost con-
straints. In the case of low-value-added granulars, technological advances must
be quite substantial before industry will commit itself to investing in an expensive
new process.
Problems are particularly varied in the light chemical synthesis and pharmaceu-
tical industries. We will mention a few typical examples.
1.2 Economic Implications and Industrial Problems 9

Powders and various granular materials are ubiquitous in pharmacology. The


specifications typical of that industry are wide-ranging. For instance, the objective
may be to mix several ingredients with different and/or complementary effects in
such a way that they can be readily absorbed with as pleasant a flavor as possible.
Such stipulations require highly sophisticated preparation techniques.
One of the best-known and simplest examples is aspirin. Its basic component-
acetylsalicylic acid-is bitter to the taste and tends to cause stomach irritation.
This product, whose patent is now in the public domain, has undergone numerous
improvements over the years, which are intended to minimize undesirable side
effects and enhance absorption into the blood stream. The starting crystallized
powder is recompacted, coated with polysaccharides (for taste), mixed with starch
(for rapid dispersion in water), and bonded to vitamin C, which gives it a yellow
color on one side and white on the other (this solution gets around the difficulty
of mixing the constituent powders intimately).
Another well-known example is the production of talcum and other cosmetic
products, which are typically composed of the appropriate mixtures of powders and
other granulars. Here again, mixing powders by convergent flows can temporarily
obstruct conduits. Such problems can play havoc with the composition of the
mixtures, with potentially disastrous consequences for the safety of the product,
particularly in the case of pharmaceuticals.
Cosmetic and pharmaceutical powders, whether of organic or mineral origin,
often go through a process of fritting, either cold or hot (such as in the fabrication
of ceramics), bonding, or even extrusion. The physics of these processes is poorly
understood and the thermal and mechanical properties of the starting materials are
rarely known with any precision. The details are much beyond the scope of this
book, but we should be aware that a number of applied physics laboratories are
trying to solve these problems of great practical interest. We will simply mention
as an example the manufacture of modem magnets intended for high-performance
applications, such as electric motors for high-speed trains. They are produced
by fritting powders of rare-earth compounds whose granulometry ranges from
20 to 50 {Lm. This high-technology industry consumes several hundred tons of
high-purity granular materials per year.
Another important application of materials in powder form, or small-size granu-
lars in general, has to do with the physical and chemical properties of their surface.
Specifically, their large surface-to-volume ratio promotes their efficient interaction
with ambient gases or liquids, which gives them a distinct advantages as catalysts
for chemical reactions. A straightforward calculation illustrates the potential ben-
efits. Suppose we start with a single spherical grain of radius R. Its volume is
~ n R 3 , and its surface area S is 4n R 2 . If we manage to divide this single grain into
N spherical particles of radius r, we obviously have N ~ (R/ r)3. The total sur-
face area ofthe N small particles is now equal to 4Nnr 2 = 4N 1/3 n R 2 = SN 1/ 3 ,
which increases as the cube root of the number of particles. An increase of the
surface area by a factor of 1000 results from 27 subdivisions, which can easily be
achieved by mechanical crushing.
10 1. Introduction

The Agriculture Industry


As is well known, many Western developed nations are major producers of agri-
cultural and food products that often rank first on their list of exports. Despite
years of steady growth, the industry has been unable to bring much improvement
to its basic processing technology, particularly of granular materials such as wheat,
com, powdered milk, cocoa, and the like. A number of research centers have been
created around the world in the hope of changing this state of affairs. Engineers
have become increasingly aware of some of the outstanding problems concerning
the processing, transport, and storage of all manner of granulars.
As is the case in the chemical industry, the processing of foodstuffs falls in
one of two categories, depending on the degree of value-added materials. The
prototype in the low-value-added category is the animal feed industry, which uses
a relatively primitive technology to handle enormous quantities of granular ma-
terials derived from agriculture. A typical plant treats hundreds of thousands of
tons of various grains per year. A small portion of its budget-usually around
9%-is earmarked for processing, while 83% covers the raw materials (mostly of
agricultural origin), and 8% is devoted to transport. It is not uncommon to notice
in such facilities hoppers pockmarked with numerous dents left by hammer blows
intended to free flow blockages caused by arch effects. Materials such as flour are
prone to creating solid plugs. It is a fact of life that these blockages have to be
routinely cleared with pick axes, if not jackhammers.
A quick survey of the problems affecting daily operations in such plants reveals
that they are almost exclusively of two types-blockages of flow by the arch effect,
and granular separation.

1.2.2. Flow Problems


Granular flows have properties that give rise to numerous problems in industrial
settings. Much can be learned by observing the flow of grains in an ordinary
hourglass filled with sand [8]. Contrary to what happens in sealed-off hourglasses
of the type once used to keep track of time, the flow of sand in cone-shaped funnels
is generally not smooth. We know from experience that the coupling between the
granular flow and the ambient fluid can cause intermittent events that impart to
the exiting flux of material a discontinuous character. How granular materials
flow down industrial-size hoppers turns out to be a rather tricky problem that has
been the subject of a great many studies. Yet the physics of this seemingly simple
phenomenon remains poorly understood.
Virtually all industrial or laboratory facilities designed to handle granular ma-
terials such as grains (food industry) or gravels (construction industry and public
works) are plagued by blocked flows which can be extremely disruptive. Block-
ages of this type occur with annoying regularity, for instance, while trucks or barges
are being loaded with gravel through giant chutes designed to handle tens of tons
of material, as schematically illustrated in Figure 4. The frequency of these events
depends on the diameter of the opening, as well as on the size and mechanical prop-
erties of the aggregates. They can invariably be traced to the formation of arches
1.2 Economic Implications and Industrial Problems 11

(b) D
(e) f\
FIGURE 4. Diagrams (a) and (b) show flow blockages due to the formation of arches near
the orifice of an hourglass or a hopper. Diagram (c) depicts an arch structure with maximum
stability (inverted chain).

in the vicinity of the discharge orifice. Depending on the circumstances, the flow
can either become sporadic or be completely blocked by remarkably stable arches.
For the same reason, it is extremely difficult to reproducibly mix two granular
materials by merging two separate flows into a common discharge channel. That
has proven to be a major problem in the food industry and in chemical plants
producing polymers, where raw materials are typically dry granulars. Blockages
can form spontaneously as a result of some kind of seizing mechanism. 9 The
phenomenon is even known to have caused dangerous explosions by allowing
fermentation gases to accumulate in grain silos plugged up by such arches.
Given the many uncertainties surrounding the static properties of a granular
system, it would seem that describing the mechanism of arch formation would be
quite a daunting problem. Nevertheless, it is possible to state a number of gen-
eral principles which will later prove useful. It is important to realize that arches,
which are a direct consequence of gravitational forces, can form because of purely
geometrical and steric reasons. This can be easily understood by considering a
stack of identical frictionless spheres resting on either side against inclined walls,
as depicted in Figure 4(b). In this example, each sphere is kept in equilibrium at
two contact points whose connecting line lies below the center of gravity. While
such arrangements can obviously form spontaneously in two-dimensional systems,
they would seem rather more improbable in real-life three-dimensional situations,
because they require the cooperation of a much larger number of spheres, each sta-
bilizing its neighbors. Nevertheless, they are frequently encountered in industrial
or laboratory conditions. Figure 5 illustrates some examples actually observed in
simple experiments.
Not surprisingly, the phenomenon of dry friction greatly increases the likelihood
of forming arches anchored onto the side walls. In this case, it is not even necessary

9The phenomenon of seizing is frequently observed in the industry, although it is not clearly understood.
It manifests itself in a gradual hardening of arches forming in hoppers. It is not uncommon for workers
in the agriculture industry to have to use pick axes to remove solid plugs in silos that have remained
idle for some time. These plugs can be the result of high humidity (see Section 2.1) or of poorly
known physical and chemical processes causing grains to stick together during storage.
12 1. Introduction

FIGURE 5. Stable arches form readily when a 1 em diameter test tube filled with fine sand
(100 !Lm) is turned upside down. The photo on the left provides a side view, while the one
on the right is taken from below to reveal the arch pinned against the wall.

for the side walls to be slanted. Indeed, we will run into several examples of arches
pinned against vertical walls (e.g., Section 3.2.4), particularly in silos. Effects of
this type can have dangerous consequences in industrial facilities. Engineers have
come up with a number of more or less effective solutions, such as those illustrated
in Figure 6.

1.2.3. Problems of Segregation


The industry processes enormous quantities of granular materials year in and year
out. Virtually every stage of its operations is at the mercy of segregation-a most
irksome phenomenon that tends to separate the components of a mixture supposed

(a) (b) (e)

FIGURE 6. Three methods used in industry to prevent or remedy blockages by arch effect.
They include (a) an Archimedes screw, and (b) a conveyor belt with a corrugated surface.
In (c), a plant worker pounds an obstructed hopper with a sledge hammer to get the flow
started again. The latter is the method of choice in industries producing low-value-added
granular materials.
1.2 Economic Implications and Industrial Problems 13

Mixture

FIGURE 7. Two techniques used in industry to try to avoid problems of granular segrega-
tion. The principle is the same in both cases. The idea is to force the materials through a
section of hardware that promotes mixing.

to be as homogeneous as possible. It is a perennial problem, for instance, in


the polymer industry, which often relies on solid phase reactions between dry
granulars that must first be mixed or fused as evenly as possible. It is no less of
a challenge in the agriculture industry, which tries to produce carefully controlled
mixtures of various grains and other proteins in granular fornl. Industries that deal
with low-value-added materials, such as animal feed, cannot afford sophisticated
technologies that would add excessive cost to their products. They rely mostly on
relative crude and not very effective techniques, of the type illustrated in Figure 7.
At the other end of the spectrum, the example we are about to discuss illustrates
the difficulties faced by industries such as plastics or pharmaceuticals. It will give
us an idea of the costly investments required when it is absolutely essential to
achieve intimate mixing of different granular materials.
The problem boils down to this: Two granular materials A and B must be mixed
in equal ratios and as uniformly as possible. The criterion for homogeneity is that
every particle of A be in contact with at least one particle of B in order to ensure
the proper reaction or fusion between the two components.
Simply dumping equal quantities of the two products in an ordinary mixer will
not do since, as we have pointed out, shaking the resulting mixture inescapably
tends to separate the two components. The technique currently favored both in
industry and laboratories turns out to be costly, labor-intensive, slow, and-least
attractive of all-difficult to adapt to mass production. But it is perfectly adequate
to manufacture modest quantities of materials. Rather than trying to prepare a
homogeneous mixture-which is virtually impossible to achieve-the solution is
to make a sort-layered cake, with alternating sheets of A and B. The trick is to start
with what amounts to two separate sheets of pie dough, one loaded with particles
of the material A, and the other with particles of B. These two sheets are laid on
top of each other, as depicted in Figure 8(a).
The sequence of steps illustrated in the rest of Figure 8 is referred to as "the
pastry-maker technique." The initial AlB matrix is stretched in one direction so
as to double its length. It is then cut in two (Figure 8(b)), and the two separate
14 1. Introduction

(a) 1
B

(b) 2 <~~
B

(e) 3 A
B

FIGURE 8. Illustration of the "pastry-maker" technique to mix two different granular


materials (after [115]).

pieces are placed on top of each other (Figure S(c», resulting in an AlB I AI B
structure. The process is repeated until the thickness of each individual layer is
roughly equal to that of the largest of the two particles A or B. At the conclusion
of N operations, the number of layers is obviously 2N . Not only is the process
extremely labor-intensive, but the final product is only a superposition of different
layers, rather than a true three-dimensional mixture.
The process just described has been adapted to mass-production with so-called
Kenics mixers, illustrated in Figure 9, which accomplish, automatically and con-
tinuously, the discrete sequence of steps of the pastry-maker technique. 10
The pastes or fluids to be mixed are injected at points A and B shown in the figure.
The ultimate speed of the process is limited by the onset of turbulence phenomena
which may appear within the reactor and can potentially upset the quality and
smoothness of the mixing. Such phenomena depend primarily on the dimensions
of the apparatus and the properties of the products to be mixed.
It is clear that implementing this technology on the factory floor requires a
large capital investment and has to contend with difficult processing techniques
(the preparation of pastes can itself be a challenge). As such, it is largely incom-
patible with the processing of industrial quantities of low-value-added materials,
whose cost has to be kept as low as possible. A better understanding of the mech-
anism of segregation in granular materials is of great importance not only from a
fundamental physics point of view but for economic reasons as well.

1.3. Granular Materials and Geophysics


To state that granular materials are ubiquitous in nature is practically a cliche.
Sand is present in enormous quantities in deserts, which happen to cover more
than 10% of the land surface of the globe. I I The shore line of oceans and lakes,

lOMixers working on this same principle manufacture the complex powder mixtures packed in rocket
boosters used to launch satellites into orbit.
11 Thestudy of deserts and desertification is a discipline with considerable impact on human affairs. It
is significant that Bagnold, one of the pioneers in the physics of granular materials, has devoted an
entire book to the topic of desert dunes [9].
1.3 Granular Materials and Geophysics 15

B
FIGURE 9. The first two elements of a Kenics mixer (after [115]).

as well as ancient sea bottoms, constitute virtually inexhaustible natural sources


of silicate sands, one of the basic materials in our civilization.
Numerous natural phenomena depend to varying degrees on the type of physics
we are about to describe. If nothing else, most involve some basic mechanisms,
such as collisions and friction, that govern the behavior of dry granular media.
By and large, geophysics concerns itself with much more complex objects than
those of interest to us here. For instance, the physics of moving sand-made of
particles in suspension in a liquid-falls in the realm of fluidized beds, which is
rather different from the types of problems addressed in this book.
On the other hand, an avalanche of snow in a state of supersaturation may be
understood in terms of concepts that will be discussed in Chapter 4 (although not all
the experts agree on this point). A great deal of caution is warranted in these areas,
even though models are largely heuristic and, as such, ought to be applicable in
principle to a variety of avalanche phenomena, regardless of their specific nature.
In the same spirit, there exists a certain analogy between the phenomenon of
granular separation by shearing, to be addressed in Chapter 5, and that of stratified
sedimentation in rivers churning rocks or sand. 12 This type of stratification occurs
in nature on massive scales, covering the entire length of a river or the height of
a mountain. This explains at least in part how rivers shift course over periods of
centuries, how they become clogged with sand, and other natural phenomena with
great ecological consequences. Likewise, the friction experienced by rocks subject
to high winds or falling down steep slopes contributes directly to the erosion of our
planet's surface and has a profound effect on the evolution of our environment. It
would obviously be much to our advantage to be able to predict the overall behavior
of all these natural phenomena. To a modest extent, the principles discussed in this
book may well bring us one step closer to this ambitious goal. Indeed, the material
covered in subsequent chapters can be viewed as a description from first principles
of various mechanisms at play in our geophysical environment.

12Guy Berthault has made an excellent documentary movie about this phenomenon, based on a study
conducted at the University of Colorado [10].
16 1. Introduction

Another discipline that, in many respects, has much in common with the physics
of dry granular materials is seismology. How forces of friction are stored and sud-
denly released is just as much of interest to geophysicists as it is to fundamental
theorists trying to understand the basic properties of granulars. The collective
behavior of particles in quasi-permanent contact is the central focus of the geo-
physical sciences. Granted, geophysicists typically deal with interactions that are
far more complex than those to which we will deliberately restrict our attention.
They have to worry about phenomena such as cohesion effects, contact fatigue,
erosion, strain hardening, and many others that we shall barely touch upon. Yet,
the phenomenon of fracturing, which is at the basis of geophysics, shows up just
as faithfully, albeit in an elementary form, in the most grossly simplified versions
of the physics of granulars, as we will see in Chapter 3.
Avalanches and free flows are treated in Chapter 4, in which we draw liberally
from the work of geophysicists. We do so for at least two reasons. First, an
accurate description of what is called the stick-slip phenomenon requires a clear
understanding of the interdependence between friction and velocity, which also
happens to be at the heart of the physics of granular materials. Second, the simu-
lation of so-called self-organized critical (SOC) systems has been called upon to
explain both granular avalanches and the scaling laws of the distribution of earth-
quakes (typified by the Gutenberg-Richter law). The validity of this approach
remains a topic of lively debate in both cases.

1.4. A Brief Historical Review


In the past, the physics of granular materials has received far less attention on the
part of researchers than, say, hydrodynamics. Yet, it is remarkable and, in a sense,
admirable that a few notable scientists managed in those days to marvel at some
fascinating aspects of the behavior of these types of solids.
The first recorded mention of granular flows is due to Lucretius (ca. 98-55
B.C.), the famous poet and natural philosopher in ancient Rome. He wrote in 55
B.C.: "One can scoop up poppy seeds with a ladle as easily as if they were water
and, when dipping the ladle, the seeds flow in a continuous stream.,,13
The scholars of the Renaissance had wide-ranging interests. Leonardo da Vinci
(1452-1519) was the first to devise a simple and convincing experiment demon-
strating the laws of dry friction. He and others were even able to make a few
pertinent statements concerning piles of sand. It was not until the end of the eigh-
teenth century, though, that Charles de Coulomb (1736-1806) wrote a definitive
paper, which is still frequently cited, entitled "Essay on the rules of Maximis and
Minimis applied to some problems of equilibrium related to architecture" [11]. The
paper in question, which is of interest in several respects, is based on a number of

13The poppy seed has become popular again of late as the material of choice to study the properties
of granulars. Since it contains water, it can be imaged by nuclear magnetic resonance (NMR) to
generate three-dimensional pictures of the collective behavior of a collection of grains, as we will
explain in Chapter 3.
1.4 A Brief Historical Review 17

experimental observations on the equilibrium of earthen embankments, the stabil-


ity of stone structures, and other edifices. It puts the physics of granular materials
on a foundation that is difficult to contest even today. For instance, it ultimately
led to the celebrated Coulomb laws of dry friction between solids, which in time
would be extended to granular materials. In that sense, Coulomb's seminal paper
can be viewed as heralding an entirely new discipline.
In 1780, Ernst Chladni (1756-1827) noticed a number of interesting differences
(which remain rather puzzling even today) between the behavior of light granulars
(such as horse hair) and that of coarser and heavier ones (such as sand). He observed
what came to be known as Chladni's complementary figures, which we will discuss
in Chapter 2. His experiments were duplicated and confirmed a short while later by
Christiaan Oerstedt (1777-1851), who used lycopodium powder-an extremely
fine material that was instrumental in many other discoveries. 14 Felix Savart
(1791-1841) was interested, among other things, in music. He would make use
in 1827 of Chladni's geometric figures to study the frequencies and wavelengths
of sound waves.
Michael Faraday (1791-1867) was another illustrious figure who, against the
backdrop of his primary work on hydrodynamic instabilities, also had a strong in-
terest in how vibrations induce the formation of sand piles [12]. This phenomenon,
closely related to Chladni's experiments, remained a mystery until recently. The
problem was to figure out the role of air as sand collects in regular patterns. Chap-
ter 3 will provide an opportunity to discuss the problem of granular instabilities,
reminiscent of those observed in liquids by Faraday, as well as the formation of
sand piles.
William Rankine (1820-1872) examined the theoretical implications of friction
in granular materials as early as 1857 [13]. Starting from Coulomb's ideas, he
established a number of principles which remain fully valid to this day. He defined
what is now called Rankine's passive and active states. These concepts are clearly
and thoroughly discussed in the book by Brown and Richards, and we will not
dwell on them here [5].
The problem of the equilibrium distribution of forces in a granular medium
stored in a silo has been studied by several researchers, who have published the
results of their work. 15 In 1884, I. Roberts noted that "in a structure whose lateral
walls are parallel, the pressure exerted by the stored grain onto the base ceases
to increase when the structure is filled to a height more than twice the diameter
of the inscribed circle" [14]. A few years later, H. Janssen, a German engineer
from the city of Bremen, proposed a model based on a coefficient describing the
redirection offorces toward the wall [15]. He failed to refer to the work of Roberts,
presumably because he was unaware of it. We will present the model in Chapter 3,

140ur great ancestors in science deserve much admiration for their inquisitiveness. Their curiosity was
rarely restricted to a single discipline.
15This problem remains quite current. Given the various uncertainties concerning the forces at play in
a granular medium-which we will study in some detail-the correct way to model the distribution
of forces on the base of a silo remains in dispute.
18 1. Introduction

and even generalize it to the dynamics of granular systems. Substantially the


same idea inspired the work of John Strutt (1842-1919), published in 1906 [16].
He mentioned Roberts' observation but not Janssen's work-again because he
probably was not familiar with it. Incidentally, Strutt (who is perhaps better
known as Lord Rayleigh, winner of the 1904 Nobel Prize in physics) suggested
an interesting analogy between this problem and the way a rope wrapped around
a pole resits traction.
By the latter part of the nineteenth century, Osborne Reynolds (1842-1912) had
already distinguished himself in the field of hydrodynamics (Reynolds's number)
[17]. He also made some fundamental contributions to the theory of granulars
around the year 1885. Some concepts he developed (notably that of dilatancy,
which we will discuss in Chapter 3) and his analysis of slanted embankments
remain high on the list of modern topics of investigation.
The number of scientists and engineers who have devoted their talents to this
discipline has grown steadily during the twentieth century, particularly since the
1950s. One individual deserves to be singled out. His name is Ralph A. Bagnold
[18]. Between 1940 and 1970, he made many important observations and wrote
a book on desert sands that has become a classic [9]. Since that time, the number
of scientific publications in that field has mushroomed. Indeed, this proliferation
was one of the main incentives for writing this book, as we have indicated in the
Preface.

1.5. Prerequisites and Selected Bibliography


Before tackling the remainder of this book, the reader is encouraged to familiarize
himself (or herself) with a few basic or review articles, which provide a bird's
eye view of the problems related to the "physics of sand piles." Among others,
we strongly recommend references [19]-[22]. For the physics and mechanics
of granular materials, the text by Brown and Richards is widely considered the
ultimate reference, and rightly so, even though it is unfortunately out of print
[5]. The texts by Bagnold [9] and Nedderman [23] are also recommended. Some
more recent works, published in the format of proceedings of various specialized
symposia, are also useful sources [6], [24]-[27]. They offer in-depth analyses of
some of the issues outlined in this book.
As stated in the Preface, the goal of this text is to present concepts and results
that are as recent as possible. In that spirit, we assume that the reader is familiar
with fundamental concepts (such as the Mohr-Coulomb model, the Rankine states,
the method of characteristics, and so forth). They are treated in detail in the book
by Brown and Richards or in the text by Savage. The kinetic theories developed
in the 1980s would by themselves justify an entire book the size of this one.
They concern themselves with relatively dilute particulates in rapid motion. This
is quite different from concentrated granular media, the topic which the present
text is entirely devoted to. Readers interested specifically in kinetic theories are
referred to other basic sources (particularly references [28], [29], [30]).
2
Interactions in Granular Media

2.1. A Single Particle and Its Environment


The physics of granular materials concerns itself with real-life situations in which
solid particles typically find themselves in a gas environment (usually air) or in
a liquid. The properties of granulars depend primarily on the nature of the inter-
actions between the particles themselves, but also between the particles and their
environment. When the influence of the environment can be neglected, we deal
with what is called dry granular materials. Otherwise, the resulting systems can be
quite complex and exhibit a variety of behaviors depending on the media involved.
This field of study, which has considerable industrial applications, encompasses
a broad range of media, from pastes to liquid muds to porous materials. It has
been the object of a large number of studies. While this topic is in the midst of
rapid development, it remains largely outside the scope of this book, which focuses
deliberately on a highly simplified approach in which interactions are reduced to
a minimum, as opposed to the much more complicated interactions characteristic
of "moist" granular materials. In what follows, we shall concern ourselves almost
exclusively with dry granulars. These have not always commanded much in the
way of sustained interest on the part of the physics research community, in spite
of the many different possible behaviors and fundamental nature of the problems
involved, which would normally be expected to pique the curiosity of physicists
and industrial engineers.
It should be borne in mind that dry granular materials, in which interactions
are simplified and reduced to friction and collisions between particles or with the
container walls, are something of a "theoretical" species, as it were. It would seem

J. Duran, Sands, Powders, and Grains


© Springer-Verlag New York, Inc. 2000
20 2. Interactions in Granular Media

that such idealized objects might be fairly accurately approximated in the labora-
tory by placing solid particles in a vacuum chamber. In practice though, things are
rarely that simple because of a major intrinsic problem. As it turns out, the energy
dissipated through mutual collisions or friction is inevitably accompanied by the
appearance of surface electrical charges on both the particles themselves and the
walls of the container. Such surface charges are particularly difficult to eliminate
in vacuum; the resulting medium-range electrostatic interactions can greatly com-
plicate the analysis of experimental results. Adding a surrounding fluid makes it
much easier to drain away those pesky surface charges. For that reason, such a
fluid is not only desirable, but indeed essential. Fortunately, provided a few pre-
cautions are observed, a fluid matrix does not necessarily preclude treating many
laboratory, or even industrial, situations involving granulars as though the mate-
rials were truly dry. The purpose of this section is to define the range of validity
of the methodologies which we will develop throughout this book. For strictly
pedagogical reasons, we start out by discussing the interactions of a single particle
with its fluid environment.

Laminar Drag
Consider a spherical particle of radius R, mass m, and volumetric density Pb,
moving at a velocity v in a fluid of viscosity rJ. The surrounding fluid is contained
within walls close to the central sphere, as shown in Figure lO(a). We wish to
compare the kinetic energy of the particle with the energy loss due to the work
done by the viscous forces existing within the fluid-forces which tend to slow
down any movement.
To that end, we define a parameter R1 (the subscript l stands for laminar), which
is simply a measure of the ratio of the kinetic energy of the particle and the work
done by the viscous force over a characteristic length, which we take to be the
radius R of the sphere. This parameter will enable us to estimate the importance
of viscous dissipation in relation to the kinetic energy. Based on what every student

Laminar Turbulent
FIGURE 10. A sphere moving in a fluid in a laminar or turbulent regime.
2.1 A Single Particle and Its Environment 21

has learned in fluid dynamics theory, we write


mv 2
R[ - -,,----------,------
- 2 JI rJlT R2(avjax) dx'
where the integral in the denominator extends over the interstitial space I around
the particle. It is not essential to work out this integral exactly. In order to get an
order of magnitude for R[, all we need to do is estimate this integral by considering
that all the lengths entering into it are of the order of R. In that case, we find that
the denominator is of the order of lOJTr]R 2 v, while the numerator is equal to
1PbJT R 3 v 2 . Within these approximations, we find that
Pb
R[ ~ -Rv. (2-1)
r]
We might point out in passing that the problem could have been solved directly
by using Stokes's formula for the drag of a sphere. We also note that (2-1) bears a
striking similarity to the standard expression for Reynolds's number, which gives
the ratio of convective flux to diffusive flux in a viscous fluid in motion. The reader
will recall that Reynolds's number is given by
UL P
Re =- = -UL,
v r]
where U is the velocity of the moving fluid, and L is a length characteristic of that
fluid. It is clear that U and L playa role similar to v and R, respectively, in (2-1).
The only difference between the two definitions is that in the case of a particle in
motion in a fluid, it is the volumetric density of the particle that matters, rather
than that of the moving fluid itself. With this restriction in mind, the number R[
introduced above can be thought of simply as the Reynolds number for the problem
at hand.
It is useful to clarify the practical implications of (2-1) with the help of a nu-
merical calculation that will put some orders of magnitude in perspective. If we
consider a particle of silicate sand, of density 2.2 g/cm 3 and diameter 1 mm, mov-
ing at a speed of 1 cm/s in dry air, R[ turns out to be roughly equal to 104 . This is a
reassuring result which gives us confidence that it is perfectly legitimate to regard
particles with this particular mass and speed as a dry granular. If, on the other hand,
we were to deal with a powder made of particles whose typical size is only 10 !Lm
(e.g., lycopodium powder), rather than the 1 mm used in the previous example,
the number R[ would decrease proportionately. It could easily get smaller than 10,
which would sound a warning that the viscosity of the surrounding fluid can no
longer be safely ignored. This simple calculation illustrates that, even in favorable
experimental circumstances (such as a noncritical system of particles in motion),
it is advisable to exercise caution. It is imperative to be alert to the possibility that
our approximations for a dry granular medium may break down when interactions
with the surrounding fluid are no longer negligible. 1

1Some recent experiments, whose significance will become more transparent after reading Chapter 3,
have clarified some of the concepts introduced here [31].
22 2. Interactions in Granular Media

Before leaving this subject, the reader will appreciate that the situation is likely
to be problematic in a fluid such as water, which has a coefficient of viscosity a
hundred times larger than air. Even with particles as small as I mm in diameter,
viscous interactions cannot be neglected in this case.

Turbulent Drag
Hydrodynamics taught us that, as soon as the relative speed of a particle in motion
in a fluid matrix becomes sufficiently high, instabilities appear in the boundary
layer between the fluid and the object in motion (Figure lOb). There is at that
moment a transition from a laminar flow to one that is turbulent. We also know
that the transition between the two regimes occurs for relative speeds of the order
of a few cmfs in dry air. As it happens, most granular flows involve velocities that
are precisely in that range. We can thus anticipate that, depending on the details of
the circumstances, energy dissipation will be either laminar (directly linked to the
viscosity of the fluid) or turbulent (related to the difference in dynamic pressure
ahead of and behind the moving object).
Here again, we define a number R t (where the subscript t now stands for tur-
bulent) as the ratio of the kinetic energy to the work done by the opposing force
of turbulence F t over the applicable characteristic length, which, as before, is
typically of the order of the radius R of the particle.
This time around, though, the force Ft is proportional to the frontal cross-
sectional area S of the particle, to the square of the relative speed v, and to a
coefficient kt whose value is typically 0.24 for a sphere. Specifically, we write
2
Pov
Ft =kt - -S,
2
where Po is the density of the fluid. We then find that the number R t is approxi-
mately equal to

1 Pb
Rt ;::::::--.
kt Po
This equation tells us that the number R t does not depend on the velocity (assuming
that the regime is indeed turbulent). It also shows that the concept of dry granular,
i.e., of a system of particles whose interactions with the surrounding fluid are
negligible, is valid for sufficiently heavy particles such as grains of silicate sand
moving through a gas such as air. 2 In this case the number R t is greater than

2The cmcial importance of the density of the medium in the physics of granulars was vividly illustrated
by a simple observation due to Chladni in the eighteenth century. He had noticed that small pieces of
horse hair torn loose from a bow mbbing against the strings of a violin tended to collect on the wooden
body of the instmment according to specific patterns. He proceeded to carry out a series of simple
experiments by depositing on a violin either fine sand or lycopodium powder. The sand piled up at
the nodes of vibration, while the lycopodium powder gathered at the vibration peaks, giving rise to
complementary figures (known as Chladni's figures). The explanation is that the very light lycopodium
powder responds to the vibrations of the surrounding air, while the sand is sensitive primarily to the
vibrations of the instmment itself. This, incidentally, constitutes a magnificent example of a physicist's
sense of observation, as reported by Michael Faraday in 1830.
2.1 A Single Particle and Its Environment 23

10 3 , which justifies neglecting the effects of turbulent dissipation. Once again,


we come to the conclusion that these effects can definitely not be ignored when
the fluid is a liquid, in which case the ratio of the densities is about unity and
hydrodynamic dissipation becomes predominant.

Granular Dendrites
We proceed next to discuss a rather spectacular and poorly understood phe-
nomenon, which is due to an intriguing type of interaction between air and granular
matter [26], [32]. Observed in granular dendrite experiments, it involves effects
that have well-known counterparts in hydrodynamics when one fluid is injected
into another more viscous one. Even the experimental configuration tradition-
ally used with liquids (Hele-Shaw radial cell) is essentially duplicated to study
granulars.
The actual experiment goes as follows: A layer of fairly loose fine sand is placed
between two glass plates. The upper plate has a hole in its center through which a
hose is inserted to blow pressurized air into the granular material. When both the
pressure and the flow are low, nothing happens; the sand behaves as an ordinary
porous medium. But as the flow is increased, the grains of sands start moving
around and a dendritic pattern develops with a fractal structure reminiscent of
what is observed with non-Newtonian fluids. An example is shown in Figure 11.
The outcome of this experiment remains something of a mystery. Many ques-
tions come to mind. How can the sand grains move at all when they are confined
in such a tight space between the glass plates, particularly since they must first
expand before they can deform, as we will learn shortly? In other words, is the
stack compact in the sense of the principle of dilatancy (see Section 3.1.3)? What
determines the size of the dendrites? Can the granular medium be considered a
viscous fluid?

Compressed-air injection tube

Shadow of injection tube

FIGURE 11. Growth of a dendritic pattern when air is injected into a confined granular
medium. The photograph on the left shows the beginning of the process. The one on the
right shows a well-developed pattern roughly 10 cm in diameter (after [26]).
24 2. Interactions in Granular Media

Humidity, Electrostatic Interactions, and Other Perturbations


There are plenty of indications, both in the laboratory and in industrial situations,
that other perturbations can complicate things and tax our ability to model the
behavior of "dry granulars" as we have defined them earlier. Ambient humidity,
for instance, can cause serious disruptions by creating clumps of particles that are
more or less mobile. We know from common experience that wet sand can be fairly
cohesive, whereas dry sand crumbles apart readily. It is also intuitively obvious
that the smaller the particles, the greater the perturbation-in the context of the
present discussion-since capillary forces are then apt to be of the same order of
magnitude as the gravitational forces that would otherwise largely determine the
behavior of these objects. Once again, some systematic simplifications enable us
to estimate how small the radius of spherical particles has to be in order for two
of them to remain stuck together by a thin layer of water, against the action of
gravity. The problem is illustrated in Figure 12.
Calculating the capillary force that keeps two wet spheres in contact is far from
trivial. Several methods have been proposed to avoid the difficulties associated
with solving the Laplace-Young equation. The simplest one is known as the "pulley
method," because the contour of the meniscus is approximated by a pulley in the
plane tangent to the two particles. The "pulley" has an inner radius rj, while
the outer radius (of the ridge) is rz [33]. This approach gives an expression for
the capillary force in the form

Fc = nYzvrz (1 + ~~),
where YZv is the surface tension at the air-liquid interface. We assume that, when

FIGURE 12. Spheres kept in contact by a liquid meniscus (after [115]).


2.1 A Single Particle and Its Environment 25

the weight of the lower spherical particle is just equal to the capillary force, the
radius of the liquid is some fraction a of the radius of the particles, or r2 ~ aR,
and the ratio r2f rl is of the order of 5. This series of approximations leads to a
result, correct within an order of magnitude, given by

R~ j4agPbYtv .
When applied to glass beads (Pb = 2.2g/cm3 ) and water as the liquid (Ytv
73 x 10- 3 N/m), the above expression suggests that the beads will remain stuck
for diameters up to 1 mm when a = 1. With a more modest amount of water
(a = 0.01), the diameter of the particles should not exceed 100 {tm. This exercise
highlights the need to control humidity levels when doing experiments with gran-
ular materials assumed to be dry. We may also conclude that industrial processes
dealing with powders (particles smaller than 100 {tm) in open air are very likely
not to conform to the type of physics presented in this book.
It is less straightforward to incorporate electrostatic interactions in a simple
model, primarily because of the difficulty in quantifying even approximately the
electric charges that develop on the surface of granulars in relative motion. It is
somewhat easier to estimate them empirically on the basis of simple experiments,
such as the following. Drop a few steel balls, 1.5 mm in diameter, into a small
plastic tube (Figure 13). Shake the tube and its contents vigorously and place it
back vertically on a table. You will then notice that a certain number of balls remain
stuck to the walls of the tube. The position of each of these balls in precarious
equilibrium obviously depends on the position of its neighbors, since they all
interact electrostatically.
Knowing the weight of each steel sphere, and assuming that it is exactly offset
by electrostatic repulsion, it is easy to calculate that each one carries an electric
charge of the order of 3 x 10-9 C (Coulomb), which corresponds approximately
to 300 {t C/kg, or a surface charge density of 4 x 10- 8 CJcm2 . These numbers
seem to agree with those usually observed in industry. Experience shows that
organic materials are much less sensitive than minerals (by about a factor of 100),
and that the amount of accumulated charge depends on the physical and chemical
nature of the surfaces. Furthermore, as mentioned earlier, humidity minimizes this
type of perturbation, although the actual mechanism is not entirely clear. Finally,
the existence of these surface charges constitutes a danger in industrial facilities,
which tend to store large quantities of dry grains (such as corn). The reason is that
they can trigger explosions of gases generated in silos by the decomposition of
organic substances. A standard way to mitigate the risk in laboratories is to coat
sensitive surfaces with antistatic sprays. This solution is unfortunately completely
impractical on an industrial scale.
Classification of Granular Materials and Definitions
It is useful to familiarize oneself with the terminology traditionally used in the field
granular materials. The following definitions are taken from Brown and Richards,
26 2. Interactions in Granular Media

FIGURE 13. Small steel spheres suspended by electrostatic interactions on the walls of a
plastic tube.

the authoritative source in such matters [5]:

• A granular material is composed of discrete solids which are in contact


most of the time. This specifically excludes objects such as fluidized beds,
suspensions, and other loose materials in granular form. A parameter known
as the fractional solid content of a granular medium is defined as the ratio of
the volumetric density of the actual granular to the true intrinsic density of
the solid constituent.
• A powder is a granular medium made of particles less than 100 /Lm in diam-
eter. It has become customary to make a further distinction between granular
powders (10 to 100 /Lm), superfine powders (1 to 10 /Lm), and hyperfine
powders (0.1 to 1 /Lm).
• A granular solid is made of granules whose size ranges from 100 to 3000 /Lm.
• A broken solid is a granular material in which most particles are larger than
3 mm. Aggregates used to make ordinary concretes fall into that category, as
do rock slides.

In light of these definitions and of our previous remarks, it should be clear that
the physics of dry granular materials is limited in practice to granular solids and
aggregates. Excluded are particulate systems such as powders, suspensions, or
fluidized beds, in which interactions with the surrounding fluid are predominant.
2.2 Interactions between Two Particles 27

2.2. Interactions between Two Particles


Only recently was it recognized that the overall properties of dry granular media de-
pend strongly on the fundamental mechanical characteristics of their constituents.
The different possible behaviors (decompaction, vibration-induced fluidization,
blockages by arching, various types of flows, all of which we will examine be-
low), under the effect of external stimuli, can be traced in large part to the nature
and importance of the micromechanical interactions of the particles, not only be-
tween themselves but also with the walls of the container. Stated differently, the
modes of local energy dissipation determine in a fundamental way how granu-
lar materials behave. That said, we must also recognize that, as important as it
may be to a detailed description of the relevant phenomena, the physics of in-
teractions between solids in relative motion is far from being completely worked
out.
We will run repeatedly into this obstacle in our efforts to describe interactions
between solids. It will be most obvious when we try to develop realistic computer
simulations of these materials (see Chapter 6). A computer may be an ideal tool
to handle N-body problems, but it does not absolve us from having to input the
correct equations describing the relevant interactions, as we will become painfully
aware of.
We now consider what happens when two objects come into contact, either
suddenly or more gradually, as happens typically in granular media. Rather than
immediately tackling the details of complex theories that are still in the process
of being refined, we will instead intentionally limit our discussion to simple but
extremely informative models, which will help us to understand two particularly
important processes-friction and collisions between solids.

2.2.1. The Laws of Friction between Solids


The "macroscopic" laws governing the static and dynamic properties of two solids
in contact are based strictly on experiments. They are remarkably simple and
have stood the test of time with amazing durability. They have survived virtually
untouched for almost five centuries, even though they describe phenomena that
have yet to be fully understood on a fundamental level. Leonardo da Vinci gets
credit for being the first to realize in the sixteenth century that, in order to get a solid
to start sliding against another, it is necessary to apply to it a tangential force that is
proportional not to the surface area in contact but, rather, to the force pressing the
two objects together. This property was rediscovered in the seventeenth century by
Guillaume d' Amontons, and verified independently by La Hire at the insistence of
the members of the French Academy of Sciences who were at first incredulous. In
1750, Leonhard Euler (1707-1783) introduced the concepts of static and dynamic
friction. All these properties are summarized in a set of three fundamental laws.
Interestingly, it was not until the second half of the twentieth century that they
received a credible explanation in terms of interactions at the microscopic level
[34].
28 2. Interactions in Granular Media

P3

~
...
:

FIGURE 14. Leonardo da Vinci's traction experiments. The force of traction T necessary
to set the disks PI, P2, and P3 is the same in both configurations.

The Three Fundamental Laws of Solid Friction


Figure 14 depicts classic experiments that lead to the following conclusions:
• The force of traction required to set the system in motion is proportional to
the total weight of the individual components

• The force of traction T is independent of the surface area of the solids in


contact. In other words, a brick will start moving with the same force T
regardless of which of its faces it rests on.
• A distinction must made between static friction, when the solids are initially
at rest, and dynamic friction, when the solids are already in motion and must
be kept in motion at a constant speed. Each case is characterized by its own
coefficient of friction-p,s for static, and P,d for dynamic. Euler demonstrated
that P,d :s: P,s·
Although no one individual can legitimately claim to have invented all three laws,
it has become customary to refer to the entire set as Coulomb's laws, because he
used them in a now famous article on the properties of granular materials that
ranks among the most frequently cited early references [11]. We too will conform
to this tradition.
We might note in passing that the coefficients of friction depend remarkably
little on the nature of materials in contact. For instance, the coefficient p, is about
1 for friction between metals, 0.7 in the case of rocks, and 0.4 for paper. A rough
explanation for this surprising result is provided by the model discussed next.

A Microscopic Explanation
Micrographs of the surface of ordinary solids such as metals reveal a ragged topog-
raphy of the type schematically indicated in Figure 15. It is crucial to understand
that Coulomb's second law, which states that the force oftraction required to set a
solid object in motion is independent of the surface area in contact with the base,
is justification enough for studying the physics of friction on the scale of a single
protrusion (since the result does not depend on the number of such features).
These ragged features, whose typical size is of the order of 1 t.tm, are so small
that, even under weak stress conditions, the deformations they experience sub-
stantially exceed the elastic regime in which Hooke's law remains valid (see some
2.2 Interactions between Two Particles 29

E::i..
1:
Ol
'Qj
I

o
o 10
Distance (flm)

FIGURE 15. Schematic magnified view of the surface of a polished metal sample.

orders of magnitude in Section 2.2.2). 3 This implies that the protrusions in contact
deform at constant pressure p until they exactly offset the normal load force N
(normal in the sense of being perpendicular to the average plane of contact, as
shown in Figure 16).
If A is the surface area in contact, then the relation N = pA holds when equi-
librium has been reached. The effect of a sufficiently large tangential force T is
to separate the surfaces in contact. There is a constant s which characterizes the
ability of the material to resist this shearing action; it is given T = sA. In this
picture, the static friction coefficient {Ls defined previously can depend only on the
physical constants of the materials in contact. Its value is

T s
{Ls = - = -
L p

As it happens, the constants sand p vary roughly in the same proportion for many
different surfaces. 4 In particular, metals are characterized by
s
0.6:::: - :::: 1.2.
p

This model accounts for the tightly clustered range of values of {Ls. It also ex-
plains the experimental observation that the contact regions between two solids
deform slowly under the weight of the upper object, which is fully consistent with
the behavior expected in the plastic regime. We shall present in Section 6.4 a
more accurate description of these laws of friction in the context of numerical
simulations.

3 Thisbecomes even more obvious if we appreciate that the real contact surface area between two solids
can be hundred to thousand times smaller than the "apparent" area. When distributed among a few
tiny spots, the stress due to weight can easily be much larger than the elastic limit.
4This statement is not true of materials whose mechanical properties are strongly anisotropic. For
instance, materials with a lamellar structure such as graphite exhibit a strong resistance to deformations
in a direction perpendicular to the stacked layers, but a weak one to shearing in the plane of these
layers. That is precisely the property that is being exploited when mechanical parts are lubricated
with graphite.
30 2. Interactions in Granular Media

FIGURE 16. Plastic deformation of contacts during friction.

Gliding and Rotations: Frustrated Rotations


The simplified stability conditions derived above apply to idealized situations.
It would be risky to transpose them without any precautions to actual piles of
granulates. A more careful scrutiny reveals a number of additional problems when
trying to model what really goes on at contact points between two particles in a
granular medium.
We will not reproduce here a detailed analysis that can be found in a number of
classical mechanics texts dealing with solids in contact. We will, however, reem-
phasize a few concepts that are necessary for an understanding of the discussion
to follow. Figure 17 highlights some important quantities needed for a complete
description of the motion of two rigid solids (S) and (Sf) around their point of
contact I.
Let Nand T be the force vectors normal and tangent to the plane P, which is
itself tangent to the solids (S) and (Sf) at point 1. 5 The curves L and V are the
loci of the contact points on each of the two solids as they rotate about each other.
The motion of any point M belonging to the solid (S) is described by a kinematic
vector equation that gives its vector speed

where Rand R f are frames of reference attached to solids (S) and (Sf), respectively.
As written, the above equation describes the motion of point M in the frame of
reference of solid (Sf). For the purpose of simplification, we assume the solid (Sf)
to be fixed. At any given instant, a generalized rotation (pivoting or rolling) is
given as a sum of two orthogonal components

W = W n + WI.
The vector WI corresponds to a rotation in the plane of the figure, while Wit describes
spinning about a vertical axis. The dynamic equation written above includes three

5The notion of a rigid solid is purely theoretical. It clearly conflicts with the picture we have painted
of solid friction, which is based on contacts at a number of points undergoing plastic defonnation.
This is only one of several difficult issues raised in making the transition from classical mechanics to
the mechanics of real solids, particularly in granular fonn.
2.2 Interactions between Two Particles 31

(8')

FIGURE 17. Geometry of two granules in contact.

different types of motion between the solids involved:

• vg = V R' (Is) is the gliding velocity of (S) with respect to (st);


• Vp = MI x W n is the angular speed of (S) with respect to (st) due to spinning;
and
• Vr = MI X Wt is the angular speed of (S) with respect to (st) due to rolling.

Rolling without Gliding


By definition, rolling without gliding implies vg = V R (Is) = O. The instantaneous
axis of rotation is a straight line passing through point I and aligned with the vector
w. It describes a motion in which particles remain in contact by rolling on, rather
than gliding past, one another. It typically applies to a string of roughened parti-
cles, acting somewhat like cogwheels, where the rotation of one member is readily
transmitted to its neighbors. A three-dimensional arrangement of such particles is
clearly a more complicated situation.
As depicted in Figure 18, the opposite rotation of two of the particles leaves
the state of the third undetermined. In the case of a compacted pile, where all the
particles are in intimate contact, it is easy to envision that rotations might in fact

Gliding without rolling

Frustrated rotation

Rolling without gliding

FIGURE 18. Possible behaviors of three granules in contact.


32 2. Interactions in Granular Media

be entirely inhibited. This phenomenon is sometimes referred to as frustration


in some other branches of physics (such as spin glasses). This condition appears
to prevail in convection movements (see Chapter 3). We will discuss these basic
concepts in more depth when we study the fragmentation of piles in Section 3.2.4.

Gliding without Rolling


This situation applies to perfectly smooth or nearly frictionless particles. However,
it is also possible to observe gliding without rolling in compact piles when rotations
are precluded for geometric reasons, in which case, w p = w r = O. Under these
conditions, the equilibrium of a string of particles simply involves Coulomb's
maximum angle given by e = tan -1 (f-ts). As depicted in Figure 18, two particles
in contact will be able to glide on each other only if their common tangent is
inclined by angle not exceeding e with respect to the horizontal. A complete
analysis requires a knowledge of the forces pressing the particles together. As
mentioned above, these forces contribute to the static equilibrium of the overall
configuration and generate tangential friction forces. Such calculations have been
carried out to study the stability of arches, a topic we will address in Section 3.1.1.
We will encounter several practical applications of this remark.

Transition from One Regime to the Other


Rotation, pivoting, or gliding of one particle in contact with another or several
others can be analyzed by writing down the general equations of mechanics of
solids in contact. In conformance with the laws of dry friction reviewed above,
whether gliding is allowed or not will be determined by the ratio of the normal
component N of the reaction force R at the point of contact between the solids to
the tangential component T along the particle's trajectory. Coulomb's maximum
angle e = tan- 1 (f-ts) defines a cone whose axis is perpendicular to the tangent
plane at the point of contact. The motion will proceed without gliding as long as
R is contained within this cone, known as the "friction cone," and without gliding
otherwise.

Stick-Slip Motion
We will deal on several occasions (see in particular Sections 3.1.1 and 4.2.2) with a
type of motion frequently observed in situations involving dry friction. The motion
in question is known as stick-slip. It typically results from interactions between
particles that are subject to the Coulomb-Euler laws of friction and also exhibits
an elastic behavior. Given that the granular materials we are specifically interested
in often conform to both requirements simultaneously, it should come as no surprise
that the stick-slip mechanism is commonly observed in granular matter. Although
we will have an opportunity to give a more general treatment of the phenomenon
later (see Section 4.2.2), we nevertheless introduce here a simplistic but instructive
description [35]. It is based on the model illustrated in Figure 19.
An object of mass m is placed on a conveyor belt moving at a speed v. The object
is also tied to a stationary post by a string of stiffness k. The friction between the
object and the belt is characterized by coefficients f-ts (static) and f-td (dynamic),
2.2 Interactions between Two Particles 33

FIGURE 19. Experiment illustrating the stick-slip phenomenon.

such that j.1,d ::'S j.1,s, in accordance with the Coulomb-Euler laws. For the sake of
simplicity, we assume that j.1,d is negligibly small.
We start at time t = 0 with the spring in a state of rest, at which point its length
is x = xo. Because of dry friction, the object starts moving at a constant speed v.
The horizontal component of the reaction associated with the friction opposes the
traction of the spring, which we write as

T = k(x - xo) = kvt.

The uniform horizontal motion continues as long as Coulomb's condition is veri-


fied, that is, as long as

T I= kvt ::'S j.1,s·


IN mg

The condition is met for a duration t ::'S t[ such that t[ = mgj.1,slkv. As soon as
t > t[, the force of friction gives way (suddenly becomes equal to 0), and the
motion of the mass m obeys the usual differential equation

d 2x
m- = -k(x - xo).
dt 2
Its has a well-known solution, which can be written in terms of a parameter Wo =
~k/m as

x(t) = Xo + A sin[woCt - tl) + a],


where A and a are integration constants obtained from boundary conditions at
time t = t[. These conditions are x - Xo = vtl and d 2x/dt 2 = v, which leads to
vtl = A sin(a),

v = Awocos(a), tan(a) = Watl, and A = vJt? + 1/w6.


The motion is sinusoidal until time t2 when the velocity of the object relative to
the belt is zero. At that instant, static friction kicks in again. The time t2 is given
by one of the roots (the one that is different from t2 = tl) of the equation

COS[WoCt2 - tl) + a] = cos(a).


34 2. Interactions in Granular Media

c
o
:E
'"
&

Time

FIGURE 20. Dynamics of a periodic stick-slip motion. Straight-line segments (stick phase)
are connected to portions of sinusoids (slip phase).

This leads to t2 = t] + 2[(n - a)/wo]. The length of the spring at that moment is

x - Xo = A sin[woCt2 - t]) + a] = -A sin(a).


Once again, the object is dragged by the belt at constant speed v until time t3. The
distance covered during the interval t3 - t2 is 2A sin(a). The complete periodic
motion xCt) is schematically depicted in Figure 20.
It is composed of straight-line segments joined by sinusoidal sections. The
period to is given by
2 2A sin(a)
to=-(n-a)+ .
Wo v
The amplitude is the same as that of the sinusoid. At low speed, the diagram
becomes a saw-tooth made of a series of straight lines. In that case, t] = mg Itslkv.
We might also point out that the value of the static friction coefficient Its can be
determined by measuring the maximum amplitude of the motion. As a further
exercise, the reader is encouraged to look at the shape of the function F (v) which
describes the friction force as a function of the speed of the belt, and to show
that it has regions of negative slope (see Sections 2.3 and 4.2.2). This simplistic
model can also explain the forced vibration of a violin string; because the bow is
normally made of hair coated with colophane, its interaction with the string is a
classic example of dry friction. In a more prosaic vein, stick-slip is responsible
for the squeaking of poorly lubricated doors, the bucking of machine tools, and
other similar phenomena.

2.2.2. Collisions and Deformations ofElastic Spheres


Frontal Elastic Collision
The frontal elastic collision of two spherical particles, illustrated in Figure 21, is
extremely simple to analyze. 6 In this particular case, the impact occurs along the
axis passing through the center of each sphere, which means that the velocities are

6The physics of collisions between solid particles is treated in detail in reference [36].
2.2 Interactions between Two Particles 35

FIGURE 21. Two elastic spheres exchange momenta during a head-on collision.

collinear. This is, of course, a highly improbable event with real granular solids,
which most often behave as inelastic objects, subject to friction, and experiencing
collisions at various angles. Nevertheless, the ideal case is useful to look at be-
cause it involves two fundamental tenets of classical mechanics-conservation of
momentum and of kinetic energy.
Referring to the notation indicated in Figure 21, the speed after impact is given by
m2 2m2
+
mj -
Uj = Vj v2.
mj +m2 mj +m2

In actuality, collisions between real granulates always entail some loss of momen-
tum and kinetic energy. Some fraction of the incident momentum is given up to the
colliding particles by exciting sound waves propagating in their interior. In the pro-
cess, a portion of the elastic energy stored in both particles is dissipated in the form
of acoustic waves or phonons that relax by heating up the mass of both particles. A
loss of kinetic energy can also result from a permanent deformation incurred during
a collision. In any event, we observe experimentally that a ball hitting a vertical
wall of infinite mass with a velocity v rebounds with a smaller velocity -cp v (with
cp :s 1). To a first approximation, and in the case of frontal collision between two
identical spheres, it is convenient to describe the velocity shortfall in the frame of

rT :n[~:l
reference of the center of gravity of the system by a matrix equation [37]

[::J ~c,t] ~ (2-2)

where c is called the "coefficient of elastic restitution." Of course, c = 0 for a


completely inelastic collision (in which all the energy is dissipated in the collision),
and c = 1 for a perfectly elastic one.
The collision of a ball with a plane of infinite mass can be described in a similar
formalism by a matrix equation

(2-3)

Such a phenomenological description ignores the details of the mechanics of the


collision, particularly the modes of energy dissipation. As such, it must be used
36 2. Interactions in Granular Media

with some caution. On the other hand, it is often sufficient to describe experimental
observations, as we ourselves will realize when we discuss numerical simulations
in Chapter 6. As long as we are prepared to adopt this approach, it is worthwhile to
emphasize once again that E is truly a measure of the momentum loss of colliding
particles. If P and pi designate the momentum of the total system immediately
before and immediately after impact, we may write in the frame of reference of
the center of mass

p = m12(vl - V2),

pi = ml2(ul - U2),

where m 12 = m 1m2/ (m 1 + m2) is the reduced mass of the system of two colliding
objects. It is straightforward to verify that (2-2) and (2-3) are indeed consistent
with the definition of E in the form
pi Ul - U2
E = -- = ----
P VI - V2

where the minus signs in front of the ratios account for the fact that the velocities
reverse direction after the collision. The change in kinetic energy llEkin after
impact is easily shown to be given by

llEkin = -~mI2(l - E2)(VI - V2)2.

Nonfrontal Elastic Collision and Rotation ofParticles


As mentioned earlier, a frontal elastic collision, that is to say, the impact of ro-
tationless particles along a path joining their centers, is an extremely improbable
event that, to a first approximation, does not involve friction. The reality is far
more complex in granular media. It includes nonfrontal collisions and frictional
effects giving rise to gliding and rolling motions that can impart angular momenta
to neighboring particles. We proceed to give a somewhat simplified treatment of
the problem [38], [39]. A more accurate description turns out to be extremely
complicated.
As an introduction to a subsequent section in which we will outline a technique
useful for numerical simulations of rotations, we develop next a classically mech-
anistic description of the gliding and rotation motions of two solids in contact [40].

A Ball Thrown Against a Wall


We consider the simple problem of the collision between a spherical ball and a
vertical wall. Let V x and v y be the components of the velocity of the center of
mass of the ball, as shown in Figure 22, at the instant to of impact. Let Wo be
its angular momentum just prior to impact. We assume that the rotation vector is
perpendicular to the plane of the figure. Under these conditions, it is intuitively
obvious that the entire problem must retain the symmetry of the plane. Our goal
is to determine the components U x and u y of the velocity immediately after the
collision, as well as the final instantaneous rotation WI. We designate by X and Y
2.2 Interactions between Two Particles 37

x
x

Y~------_-"'l<I>

FIGURE 22. Geometry of a ball bouncing off a wall.

the components of the linear momentum at impact. As indicated in Figure 22, we


necessarily must have Vx < 0, U x ::: 0, and X ::: O. If friction can be neglected in
this simple experiment, then the solution is obviously identical with that obtained
in the previous section. We simply write U x = -8V x ' The tangential component
of Y of the momentum (X, Y) is zero, and the fundamental theorems of classical
mechanics yield:

If, on the other hand, we introduce a nonzero friction between the ball and the wall
in the form of a single coefficient fL, we must distinguish two separate cases.

1. The Gliding Velocity Remains Positive for the Entire Duration of the Collision
In this case, the system of equations describing the exchange of momenta (linear
and angular) reads

m(u x - vx ) = X,
m(u y - vy) = -fLX,

~ma2(wl - wa) = afLX,

where m is the mass of the ball (assumed solidly filled), and a is its radius. It
is immediately apparent that this system of four equations with four unknowns
(X, u X , u y , and wI> can be solved if all relevant quantities before the collision are
given and if the coefficients of friction fL and elastic restitution 8 are known. We
leave this as an exercise, and focus our attention on the phenomenology of the
collision, i.e., on the gliding and rotation motions of the ball during the collision.
We note first that the normal momentum X = -mvx(l + 8) is indeed positive,
as it should, since V x is negative. The final gliding velocity must also be positive,
38 2. Interactions in Granular Media

which is the case as long as

7 V y - awo
2: fh (1 + E) < -'---
-v x

2. The Gliding Velocity Drops to Zero at any Time t1 During the Collision
In that case, the general relations read

m(u x - v x ) = X,
m(u y - vy) = Y,
~ma2(w1 - wo) = -aY,

u y - aWl = O.

This time around, we have a system of five equations in five unknowns, which can
easily be solved. We still have to verify that the linear momentum makes an angle
less than tan -1 (fh) and, in addition, that

The last two conditions turn out to be mutually exclusive. It follows that the
behavior will be of one type or the other depending on the value of the friction
coefficient fh. 7
We may now ask ourselves whether these mutually exclusively situations are the
only two possible outcomes. Is it at all possible for the gliding velocity to change
sign during the interval [to, td of the collision? If the answer is yes, the condition
IY I < fhX could still be verified without us being able to make any prediction
whatsoever about the final value of (u y - aWl). The following reasoning gives the
answer.
Consider an instant t E [to, td, and let I] and l; designate the components of the
reaction force during impact. The equations of motion at that particular instant are

dvx
m-- =1],
dt
dv y
md"t =l;,

2 2 dw
sma dt = -l;a.

7We note here that this mechanistic model assumes that the forces of friction come into play instan-
taneously during the collision. From the point of view of a physicist, this condition is unlikely to
be realized in all situations. Indeed, phenomena of friction (Section 2.2.1) and collisions between
solids (Section 2.2.2) involve plastic deformation, a process that is notorious for being not at all
instantaneous.
2.2 Interactions between Two Particles 39

(a) (b)
FIGURE 23. Two spheres (a) before and (b) after a nonfrontal collision. The parameters
are defined in the text.

These give the following conditions concerning the gliding velocity

d
= :is.
7
m dt (V y - aw)

As it turns out, the quantity t;, which is the tangential component of the reaction
force during contact, has a sign opposite to that of (v y - aw). It follows that it can
only decrease in magnitude. Should it reach zero at any time during the collision,
it will remain zero up to the end. Stated in plain language, should the gliding
speed happen to drop to zero at any time during a collision, the ball would then
go on gliding for the remainder of the collision. This simple argument shows that
the two conditions considered previously are indeed mutually exclusive and that
they are the only two possibilities. One or the other must prevail depending on the
values of the friction coefficient JL and of the components of the incident velocity.
The same distinction between two regimes will come up again when we study
nonfrontal collisions between two spherical particles.

Nonfrontal Collision Between Two Elastic Spheres with Friction


We have already introduced two fundamental parameters JLs and JLd to characterize
interactions in the case of dry friction (Coulomb's law), as well as the coefficient
8
E: which is a measure of the momentum loss. In addition to those, we now invoke
another parameter, which we call the coefficient oftangential restitution, denoted f3.
Its effect is to dampen the tangential velocity at the point of contact as the particles
recoil after colliding. The need for such a coefficient and its precise definition will
become clear as we further discuss the process of nonfrontal collisions. For now,
let us simply consider two particles of diameter d j and d2 , and masses mj and m2,
respectively. The positions of the spheres are described by vectors rj and r2. The
geometry is depicted in Figure 23. The unit vector associated with the normal to
the plane of contact is denoted n. It can be written as

n=

8We have already been introduced in Section 2.2.1 to an elementary mechanical treatment of such
problems of gliding without rolling or rolling without gliding.
40 2. Interactions in Granular Media

The relative velocity of the two particles at the point of contact is given by

where Vi and Wi are the translation and rotation velocities, respectively, of particle
i before the collision. Note that the magnitude of the relative speed IVe I increases
when the individual velocities point in opposite directions and when the rotation
vectors of these particles point in the same direction. The velocity Ve has a normal
component V~,) = (v e . n)n, while its tangential component is v~r) = Ve - V~').
The vector v~r) defines a unit tangential vector such that t = v~)/ Iv~t) I. The angle
of impact y is defined as the angle between the normal n and the relative velocity
Ve , and is such that y E [n/2, n]. Figure 23(a) shows the situation of the two
colliding particles in the particular case when WI = W2 = O.
We now consider what happens to the kinetic momentum during the collision,
using the same convention as before, namely that Oi designates the speedfollowing
the collision. We have

(2-4)

The normal component of ~p has no effect on the angular rotation speed, but the
tangential component does. By creating a torque on the arm defined by the vector
-(dJf2)n, the change in momentum ~p causes a change in angular momentum
such that
21 ,
-nx ~p= dew -w), (2-5)

where 1 is the moment of inertia with respect to the center of the particle and Wi
is the unknown angular rotation speed after the collision. It should be appreciated
that (2-5) indicates that the change in angular momentum is the same for both
particles. Figure 25(b) depicts the situation after impact.
If ~p is known, it is possible to calculate all relevant velocities following the
collision with the help of (2-4) and (2-5). The results are
~p
OJ = V j + - ,
mj

~p
02 = V2 - - ,
m2

dj
Wi = Wj - -n x ~p
1 2It '
I d?
W = W2 - -- n
x ~P.
2 2h
The reader will recall the definition of e, given by
0(11) = -ev(lI) (2-6)
e e '
2.2 Interactions between Two Particles 41

which relates the translation velocities before and after the collision. Equa-
tion (2-6) applied to the normal component of the sum ~P/ml + ~P/m2 allows
us to calculate the normal component of the change in linear momentum, giving
the result

~p(n) = -mdl + s)v~n), (2-7)

where, as before, m 12 is the reduced mass of the system composed of m 1 and m2.
Coulomb's law establishes a relation between the normal and tangential compo-
nents of ~P, namely 1~p(t) 1 = ~ I ~p(n) I. Since collisions are always dissipative,
the vector ~p(t) must be aligned along -to We can thus write

~p(t) = ~mdl + s)v c cos(y)t, (2-8)

where we have v~n) = -Vc cos(y), since cos(y) is always negative. We also have
the relation t = v~t) /[v c sin(y)]. Combining (2-7) and (2-8) yields an expression
for the change in momentum ~p

(2-9)

The Tangential Restitution Coefficient


It is interesting to consider what happens in the case of a central collision, which
corresponds to y ---'T n. Since cot(y) ---'T -00, me change in momentum ~p would
diverge to infinity, which is physically impossible. The equations we have just
derived describe all collisions, including tangential ones, on the basis of Coulomb's
law and the coefficient of elastic restitution, and nothing else. On closer scrutiny,
though, there is another important physical mechanism that has been ignored in this
description. Specifically, at the moment of impact, the two interacting particles can
either roll on each other without gliding, or the other way around. The possibility
of a pivoting motion of one object on another (see Section 2.2.1) was not included
in the simplified model presented above. Such a motion involves moments of
inertia about the center or the rotation axis of the interacting particles.
Rather than develop a more complex model to account for rolling without gliding
(or vice versa), it is more expedient to heuristically generalize the existing model by
introducing an additional parameter, which we will call the coefficient oftangential
restitution, denoted fJ. It reflects the ability of a system to break the microscopic
contact regions during a collision, thereby initiating a gliding regime. It can be
shown that, under such conditions, (2-8) must be modified to a new form [41]

(2-10)

The second term in (2-10) has the same form as the first, except for a factor of ~
which comes about through the bias of the moment of inertia of a solidly filled
sphere. 9 In light of our previous remarks, the coefficient of tangential restitution fJ

9This factor is obviously different for other shapes of colliding particles. For instance, it is equal to ~
for disks, and 2 for thin rings. -
42 2. Interactions in Granular Media

must be equal to the smallest of two values fJo or fJl corresponding to one of two
different regimes:

• Large values of the angle y (y :::: Yo) correspond to a gliding contact, in which
the contacts have been broken. In this case, the appropriate choice for fJ is
fJ = fJo, withfJo E [-1, +1] .
• Small values of y (y ::::: Yo) correspond to a collision for which the interaction
can be described in terms of dry friction and elastic restitution. In that case,
the correct choice is fJ = fJl, with fJl = -1 - ~ fL(l + E') cot(y).

The angle Yo marks the dividing line between the two regimes, which occurs when
fJo = fJl. The value of Yo is given by the equation
7 1 + E'
-tan(yo) = -fL--.
2 1 + fJo
Reference [41] gives additional details on some practical aspects of the two modes
of collision by considering the ratio of the normal and tangential collision veloci-
ties.
This all-or-nothing simplification makes it possible to pursue numerical calcu-
lations of situations involving nonfrontal collisions and rotations. We will discuss
such models using rigid spheres to simulate particles undergoing multiple colli-
sions (see Sections 3.2.4 and 6.1.2).
Before doing so, however, it is useful to first explore in more depth a number
of problems associated with the mechanics of collisions. To that end, we proceed
to calculate some orders of magnitude of a few phenomena occurring during the
frontal collision of two spheres that can penetrate each other, as is almost always
the case in the real world.

Penetration During Frontal Collision: Hertz's Problem


Consider two identical spheres of radius R and mass M, approaching each other
at a relative velocity v, as shown in Figure 24. Hertz has shown that the elastic
energy E e stored by two spheres deformed over a depth h is given by

E e -- lkh 5/ 2
2 '

FIGURE 24. Interpenetration of two spheres during a frontal collision.


2.2 Interactions between Two Particles 43

where the coefficient k is defined as

k = 4~ _E_v'R. (2-11b)
15 1 - (J2

The quantities E and (J entering in (2-11b) are Young's modulus and Poisson's
coefficient, respectively. Upon impact, the initial kinetic energy is converted partly
into a reduced kinetic energy and partly into stored elastic energy. Thus, we write

dh)2
Mv
2
= kh 5/ 2 + M ( dt (2-12)

The velocity drops to zero when the two spheres have penetrated each other by a
distance h o. At that moment, dh/dt = 0, and we have
2/5
ho - M V4/ 5 (2-13)
- ( k ) .

The entire duration r of the collision (penetration up to h o and recoil) is obtained


by integrating (2-12)

rho dh 4fir(~) (M 2)I/5


(2-14)
r=2 Jo /v 2 -(k/M)h 5 / 2 = 5r(!o) k2 v
Evaluating the gamma functions numerically leads to the result

M 2)1/5
r = 2.94 ( k 2 v

The exponent giving the dependence of r on v is only ~. Thus, the duration of


the impact between two particles is only a weak function of their initial relative
velocity. For an order of magnitude calculation of the time involved, we consider
the case of aluminum beads, 1.5 mm in diameter, moving toward each other at a
speed of 5 Cln/S. The constants relevant to aluminum are as follows

E 6 x 1011 dynes/cm 2 ,
(J 8:::i 0.3,
k 7 X 10 10 ( cgs units).

Using these numbers, the result is r = 5 x 10- 6 s.


The purist might object that this derivation is based on a static analysis, which
neglects two potentially important phenomena:
• The elasticity limit may be exceeded. If so, the deformations are really plastic.
With the help of (2-13), we may calculate the value of ho for the aluminum
beads considered above. The result is h o 8:::i 2 {Lm, which implies a relative
deformation-or strain-of 10- 3 . Aluminum is known to deform plastically
for stresses of 20 kg/mm2 , which corresponds to strains of 3 x 1O- 3 -not
44 2. Interactions in Granular Media

very much larger than the number calculated above. The beads need only
have a velocity two or three times larger to reach the plastic regime. Most
importantly, the deformation is of the order of the height of the surface rough-
ness. In accordance with what was said in Section 2.2.1, the deformation of
the surface protrusions during a collision is necessarily plastic. The impli-
cation is-and experiments confirm it-that the degree of polishing of the
interacting objects plays a crucial role in the process of restitution during
impact [34J. We will see shortly an approach to describe the penetration of
inhomogeneous spheres with a surface layer softer than the core.
• Part of the energy may be dissipated by exciting sound waves inside the solid
beads. As an example, we may calculate the time Tph it takes a phonon to
complete a round trip across a sphere, knowing the propagation velocity Vph'
The result is Tph = 2R/v ph = 0.15 cm/6 x 105 cm/s = 2.5 x 10-7 s. In other
words, there is plenty of time for several round trips during a typical collision.
This effect constitutes a potentially important perturbation that in some cases
may invalidate the model embodied in (2-12).

Inhomogeneous Spheres: The Soft Crust Model


As we have just noted, the quality and nature of the interface between two solids
can raise serious questions about the validity of Hertz's collision model [42]. In
the case of spherical particles whose surfaces are flawed or chemically altered,
for instance by oxidation, de Gennes has proposed what might be described as a
"soft crust" model [43]. The idea is to consider that the sphere is wrapped in a
thin layer of a softer material, of thickness e. Young's modulus E e of the outer
layer is assumed to be smaller than the value E prevailing in the interior. As long
as the penetration is small (h « e), it is clear that the deformation due to impact
remains confined to the outer layer of thickness e and will not affect the core of
the particle. The situation is illustrated in Figure 25.
The assumption of a thin crust leads to a modification of the exponent in Hertz's
expression relating force and penetration, which goes as F ex h 3 / 2 and applies to

(a) (b)

FIGURE 25. Schematic ofthe zone affected during the penetration of two (a) homogeneous
and (b) inhomogeneous spheres.
2.2 Interactions between Two Particles 45

homogenous spheres. In order to understand the reason for the change, it might
be worthwhile to first outline an argument-originally advanced by de Gennes-
explaining the usual ~ exponent.
Consistent with the notation of Figure 25, h is the depth of penetration during a
collision, and R is the radius of each sphere. From elementary geometry, we can
easily derive an expression for the radius a of the contact circle. Assuming that
h is very much smaller than R, which is justified if we insist on remaining within
the elastic regime, the result is a 2 ~ Rh.
It is not unreasonable to assume that the deformation of colliding spheres affects
a depth roughly equal to the radius a of the contact circle. Under these conditions,
and in the presence of a stress P, the classical definition of Young's modulus gives
a relation of the type

h
P ~E-.
a

Reananging terms and combining equations leads to

This justifies the ~ power law. If we now consider a sphere covered with a soft cmst
of thickness e much smaller than the radius R of the sphere, but sufficiently large
compared to the penetration depth h, we can assume that the deformation will be lo-
calized within the outer layer whose Young's modulus is E e « E. In that case, the
depth affected will be of the order of e, rather than a as in the previous case. In other
words, a stress P acting on the sphere now produces a penetration depth given by

The same algebraic manipulation as before now yields

The relation between penetration depth and force has changed from h ex p2/3,
valid in the case of Hertz's homogeneous spheres, to h ex pI /2. The deformation
increases more slowly with applied force. This makes sense if we note that, in
the presence of a soft crust, the deformation is more localized to regions near the
contact point and does not spread as deeply into the core.
Such a dependence can explain, at least in part, certain experimental observations
that deviate notably from what would be expected on the basis of Hertz's model
[44]. This is not the only possible explanation. Others have been proposed, invoking
notions of disorder and enrichment in the vicinity of contact points under the
influence of stress [45]. We shall discuss a particular example in Section 3.1.2.
46 2. Interactions in Granular Media

FIGURE 26. A sphere rolling down a rough surface looses kinetic energy by friction and
successive collisions. Its trajectory is a mix of ballistic arcs and sections of curves following
the rough surface.

2.3. A Single Particle on Top of a Granular Medium


Thus far, we have considered the elementary problem-albeit difficult enough
as it is-of the interaction of two particles only. We eventually need to address
more general granular materials composed of many indistinguishable particles.
To that end, we will introduce laws that are often derived from the mechanics of
continuous media, just as the kinetic theory of liquids is a stepping stone to the
Navier-Stokes law. In the meantime, we continue down the logical progression of
our discussion by considering the case of a single mobile particle interacting with
a granular medium assumed to be rigid.
As schematically indicated in Figure 26, we can expect such a particle falling
down a bumpy slope, whose protrusions are roughly the same size as the particle
itself, to follow an irregular and complex path depending on whether it remains in
contact with or bounces off its support. We can even predict that this path, made
of sections of parabolas (ballistic trajectories) and sections of curves following
the contour of the support, will depend strongly on the mechanical characteristics
(coefficients of restitution and of friction) of the materials in contact. We can
further predict that the path will depend in some complex fashion on the speed of
the particle. Under such conditions, is there any prospect of ever writing a global
equation describing, even approximately, all these possible behaviors [20]?
As a starting point, it is useful to review Newton's law, which describes the fall
of a particle of mass m and diameter D down an inclined slope making an angle
8 with the horizontal plane. The particle interacts with its environment through
the bias of a drag force F somewhat analogous to a friction. At this point, we
may introduce a normalized acceleration r, a notion we will resort to frequently
in what follows. It is a dimensionless number equal to A/g, where A is the actual
acceleration experienced by the particle and g is the acceleration due to gravity.
Newton's law can thus be written as
. F
r = sm(8) --. (2-15)
mg

A uniform motion results whenever r = 0, i.e., when the friction force exactly
offsets the effect of gravity. It then becomes evident that the problem boils down
to working out a plausible expression for the force of friction F, as we will see in
several examples later in this book (for instance, in Section 4.2.2). The idealized
2.3 A Single Particle on Top of a Granular Medium 47

case described above helps us to understand that the force of friction corresponds
to a loss of energy as the particle proceeds down the slope. This allows us to write
F = aEjax, where x corresponds to the distance covered. The energy of the
particle of mass m has two components:
(1) its kinetic energy E b which is dissipated over a characteristic distance A as
a result of successive collisions and friction with the support; and
(2) its potential energy E p' which also decreases as the particle drops down
into a succession of wells of typical depth D.
Thus we may write
E k = 'i1 mv 2
and
Ep = mgh cos(e),
where h is the altitude of the particle.
If we consider that the loss of kinetic energy occurs over a characteristic length
A (of the order of a few diameters D), we obtain the first component of the fric-
tion force Fk = m v 2j2A. In order to derive a suitable expression for the second
component F p , corresponding to the loss of potential energy, we use the following
argument:
• At high velocity, the freely moving particle executes ballistic flights whose
typical range is of the order of D. This range can be estimated independently
(!
by considering ballistic flights of height f"...h ~ g T 2 ), where T is the typical
time of free flight of a particle moving at speed v. In this approximation, the
quantity vT corresponds to a typical distance of the order of D. Under these
conditions, we have T ~ Djv, and the force of friction equivalent to the loss
of potential energy can be written as
mgf"...h 1 2D
F p = ~ ~ 'img v 2 ·

The equation just derived suggests that the corresponding dissipative force
becomes less and less effective as the speed of the particle increases, which
certainly conforms to our intuition.
• At low velocity, the path of the particle follows the profile of the surface of
the granular support. The resulting loss of potential energy should no longer
depend on speed. It corresponds only to the dry friction experienced by the
particle as it moves down the surface. In this case we have
F p = mg.
• At intermediate velocity, we simply interpolate between the two extremes
with an equation of the form
1
F ~mg--------:,------
p 1 + {3(v 2 jgD)'
48 2. Interactions in Granular Media

where fJ is treated as an adjustable parameter. This heuristic expression


converges toward the proper limits when v 0 or v 00.

In a final step, we gather both components together in a single expression of the


form

F = Fp + Fc = mg [ t )
l+b v 2 jgD
+ c (~)],
gD
(2-16)

where a, b, and c are dimensionless constants.


It is instructive to consider the implications of this very simple model. Figure 16
illustrates the trend of the drag force described by (2-16), with the parameters set
to a = b = 1 and c = 0.25. One interesting feature of our elementary model
is that friction does not vanish when v = O. Even though the system is then at
rest, it still experiences an external stimulus since the curvature of the graph points
downward. The particle is in unstable equilibrium; it will start moving when it
finds itself on a slope exceeding what is called the angle of movement 8m , in which
case the friction force is precisely equal to mg cos(8m ). The friction subsequently
decreases as the velocity increases, and reaches a minimum value mg sin(8r ). The
angle 8r corresponds to a stable equilibrium since the curvature of the graph now
points upward. It is called the angle of repose. The corresponding velocity is
equal to Vil5. To the right of that point, the effective friction force increases
rapidly, which tends to bring the operating point back down toward the minimum.
To a first approximation, the system is constrained to evolving between 8m and 8r
only.
We will have an opportunity to make use of all these concepts again when
we study avalanches in Chapter 4. But this is a good place to emphasize how
simple considerations of exchange of kinetic energy dissipated by friction and
potential energy give rise to two equilibrium states, one of which is stable and the
other only marginally so. This turns out to be a general characteristic of granular
flows.
We may also point out a connection with the stick-slip phenomenon discussed
in Section 2.2.1. As we have seen, when the velocity is zero, the system is kept in
a stationary state-although in unstable equilibrium-determined by dry friction
(coefficient f-Ls). Under appropriate conditions, it will evolve toward a state of
finite speed so as to minimize friction (coefficient f-Ld < f-Ls). Pushing this analogy
further, we can anticipate that a real system will tend to oscillate between these
two states when subjected to an external stimulus, as will indeed be confirmed
in our analysis of avalanches. Likewise, we may also predict the existence of
interesting hysteresis effects associated with the fact that a particle will evolve in
different ways depending on the initial conditions. If it starts with a finite speed,
it is likely to settle in a state corresponding to the minimum shown in Figure 27.
If, on the other hand, it is initially at rest, dry friction will tend to keep it in that
state.
More extensive studies of the fall of a sphere down an inclined plane lined with
similar fixed spheres have recently be published by several research groups. They
2.4 Interactions Between Several Particles 49

LLI~ 3 a=b=1
c
o
t5 2 c= 0.25
it
""0

co
E
o
z 0 +---+-----.-----.--~
o 2 3
Normalized Velocity ~

FIGURE 27. Effective friction force F as a function of the particle's velocity. The curve
does not go through the origin and exhibits a minimum.

attacked the problem either experimentally or through numerical simulations [46],


[47]. The reader is urged to consult these references for more details.

2.4. Interactions Between Several Particles


2.4.1. The Laws ofFriction in a Granular Medium
Having mastered the laws of friction between solids reviewed earlier in this chapter,
we are now in a position to explore how a real granular material reacts to the kind
of shearing force involved in Leonardo da Vinci's experiments. The experimental
device illustrated in Figure 28 was originally devised and used by Dawes [5].
It consists of two open boxes with individual compartments shaped like a comb. 10
The two boxes are filled with a granular medium (such as a powder or river sand)
and placed upside down on top of each other. A lateral shearing force F is applied
to the upper box, which is also subjected to a vertical load P.
Without going into the details ofthe experiments, we state the following exper-
imental observations:

• The shearing force required to set the upper half in lateral motion is strictly
proportional to the load P and is independent of the surface area under shear

F=~sP.

This result is entirely consistent with the first law of solid friction reviewed
in Section 2.2.1. 11
IOSoil-mechanics scientists routinely use boxes of this type, roughly a meter long, to measure granular
friction coefficients out in the field.
II It appears that Dawes did not try to verify that the force F was independent of the surface area of
the boxes he was using [51. It also appears that Coulomb implicitly accepted this independence by
writing that granular materials obey the laws of friction as stated by d' Amontons-La Hire [II]. This
was indeed confirmed in recent experiments.
50 2. Interactions in Granular Media

FIGURE 28. Typical experimental setup used to measure the coefficient of the static friction
fhs of a granular material.

• The coefficient fts is of the order of 0.7, which is quite comparable to values
typically encountered in the case of friction between solids. The angle e,
given by the formula e = tan-lefts), is approximately equal to 35°. It is
useful to keep this value in mind, as it will resurface in Chapter 4.
We note in passing that these results hold equally well for cohesive and nonco-
hesive granular materials (cohesive materials are made of particles kept together
by cohesive forces). The only difference is that in the cohesive case the above
expression must be modified by the addition of a constant term C; the expression
then takes on the form F = fts P + C.
These results have actually been long accepted by people working in the field of
soil mechanics, who routinely make use of the concept of stress. 12 Nevertheless,
these properties constitute a rather stunning discovery. Noone could have predicted
them, given the complexity of granular piles and the many unknowns concerning
their equilibrium (see Section 3.3.1). In particular, it seems quite unlikely that a
model based on plastic deformations, such as we invoked in our discussion of the
laws of friction between solids, could be extrapolated to the present situation. The
best we can do at this point is to accept that a medium as complex and heteroge-
neous as a granular can lead to exceedingly simple macroscopic laws. As an aside,
the same can be said of two sheets of cardboard in contact [48].

2.4.2. Bagnold's Number


We now proceed to consider the dynamics of a flowing sheet of granules at most
a few particles deep. When a container filled with such a material is tilted beyond
a certain angle, the resulting flow normally involves only a limited number of
upper layers, say, from 6 to 10. An analysis of this situation, fairly similar to
the one we will outline, was carried out by Bagnold [18]. It led to a parameter
characterizing the flow of granular materials, which is reminiscent of the familiar
Reynold's number in hydrodynamics.

12Experts in soil mechanics use Coulomb's law by assuming that the shear stress T is related to the
normal stress (}Il by T = fhs(}n. In doing so, they implicitly admit that a stress tensor can be defined
in a granular material (which is far from evident) and that, consequently, the force required to set
the object in motion does not depend on its surface area. The assumption tums out to be justified in
most situations.
2.4 Interactions Between Several Particles 51

FIGURE 29. The flow of a granular material medium normally involves only a few layers
near the free surface. The mean velocity of the particles decreases rapidly with depth.

Consider a cylindrical container filled with a granular suspension, i.e., a collec-


tion of granules in a viscous liquid. A second smaller cylinder, concentric with
the first, and dotted with large protuberances, is rotated inside the first. We are
interested in the resulting flow pattern of the particles. The following discussion
also applies to an "avalanche" flow, obtained for instance by tilting a table covered
with a thick layer of granules, as illustrated in Figure 29.
Experiments show that any flow takes place in sheets. The movement of the
granular material occurs in stratified layers parallel to each other, each layer being
characterized by its own velocity, different from that of the adjacent layers. The
medium can be viewed as a succession of layers slowed down by friction with
layers on either side. We can introduce a parameter called the rate of shearing of a
sheet of granular material in motion on top of another sheet moving at a different
velocity. This parameter is a measure of the average velocity gradient and is de-
fined by y = (avjaz), where z is the depth. It is evident that the rate of shearing
must figure prominently in constructing our model, particularly to account for the
loss of kinetic energy by friction. Let D be the diameter of individual particles
of mass m. The quantity Dy measures the relative velocity of this particle with
respect to the underlying sheet. The excess loss of kinetic energy, which occurs
over a characteristic length Ae typically of the order of a few diameters D, is due
to the friction force Fe. This we may write

- - ~ - - y2
Fe -_(aEe)~mD2 .
ax 2A e

We also note that the quantity 17Y, where 17 is the viscosity of the interstitial
fluid, represents the friction force F v due to viscous interaction with the fluid,
divided by the cross-sectional area of the solid in motion, which is of the order
of D 2

Bagnold introduced a number B defined as the ratio of the forces due to friction
52 2. Interactions in Granular Media

and collisions between solids to the forces exerted by the viscous fluid. 13 In other
words
Fc my
B=-~--
Fv 2A e 17
Two distinct regimes can be defined depending On the relative importance of Fc
and F v . In particular:
• B < 40 corresponds to a regime known as macro-viscous, in which energy
is dissipated primarily by way of viscous interaction with the ambient fluid.
Physically, the fluid does most of the work to set the particles in motion. Only
rarely do the particles come in contact with each other. Actual examples
of macro-viscous systems include muds, moist pastes, and solid particles in
suspension.
• B > 450 corresponds to an essentially granular regime, in which most of the
energy is dissipated by collisions and friction between the solid particles.
The intermediate case requires some caution.

13 Actually,
Bagnold started out by introducing the notion of dilation 8 of a granular mixture, defined as
8 = D/A e . Bagnold's number is then equal to 8 1/ 2 Ps D 2 Y/1J. This quantity merges with the number
B defined in the text in the limit of concentrated solutions, for which (Dy)I/2 is indeed of the order
of D.
3
Fluidization, Decompaction,
and Fragmentation

Common sense tells us that, in the absence of any external perturbation, a pile
of granular material is at rest. It can be inclined, gently shaken, and submitted to
various mild stimuli without anything drastic happening, in other words without
triggering a flow or collapse, and without affecting the relative positions of the
granules. We also know from experience that a sufficiently high inclination can set
off an avalanche, possibly even a continuous or intermittent flow; that vibrations
cause a granular material to behave somewhat like a liquid; and that sand deforms
readily when crushed. All these properties are markedly different from those of
the normal solids, liquids, and gases we are accustomed to. The main difference
appears to be that granular materials must be "sufficiently" perturbed in order to
change. This observation strongly suggests that we will be dealing with issues
such as thresholds, nonlinearities, and hysteresis effects.
The aim of this chapter is to clarify some principles and highlight some of the
difficulties encountered when trying to understand why and how a granular pile
changes configuration. We are particularly interested in phenomena of decom-
paction,jiuidization, and fragmentation. As the chapter progresses, we will come
to realize that these phenomena can overlap, sometimes even act in concert, and
that they constitute different facets of a more complex reality. For instance, we
will show that progressive decompaction and convection, which show up in long-
term observations, can actually result from a subtle superposition of discontinuous
processes and fragmentations that typically occur over short periods of time. What
this means is that, as is sometimes the case in other branches of physics-most
notably in hydrodynamics-the evidence for a particular phenomenon depends on
the duration of the observation.
However, before tackling the problem of how granular stacks deform, we will
take the time to analyze briefly the static properties of a sand pile. Paradoxically,

J. Duran, Sands, Powders, and Grains


© Springer-Verlag New York, Inc. 2000
54 3. Fluidization, Decompaction, and Fragmentation

...
·. . .
.. ·... ..
..
... .· ....' ....
_M"~_·_· ._. .. . _,,_,_
_~

Cannon Ball Stack Contact Points

FIGURE 30. The classic "cannon ball" stack. Even though the structure appears to have a
high degree of order, many parameters are actually undetermined.

the static properties pose even more delicate challenges than their dynamic coun-
terparts, which we will examine later.

3.1. The Static Properties of a Granular Pile


3.1.1. First Principle: The Role ofFriction
The very nature of contacts between solids and, more importantly, the mechanism
of solid friction (see Section 2.2.1) govern the static and dynamic properties of
granular media in a complex way. When dealing with systems of many particles,
it becomes quickly apparent that the problem is almost inextricable because it
has a large number of solutions that depend very much on the past history of the
material. The solution Nature chooses is almost never predictable, as it involves
some degree of disorder of the forces of contact between particles. It is essential
to clearly understand this point before proceeding any further. The goal of the
next few paragraphs is to convince the reader of the intrinsic complexity of the
problem at hand, as well as to highlight some key ideas that will prove of great
use subsequently.

The Stacking of Cannon Balls


In order to demonstrate that what appears at first sight to be an exceedingly simple
and highly ordered geometric configuration can actually be highly complex, we
consider the stack depicted in Figure 30. This structure is known as a cannon ball
stack. Assuming that the balls are perfectly smooth and absolutely identical, it
is evident that in the absence of a retaining wall the lower row has to be pinned
or glued to the ground. With this restriction, the static properties of such an
arrangement, even if it contains a vacancy (marked V in the figure), presents no
difficulty at all. The solution is well known to structural engineers.
The problem of a real stack is of a different nature altogether [49].1 There
are two main reasons for this, which introduce a degree of randomness in the
geometrical pattern of the contacts and of the force chains.

1It is
interesting to note that the static properties of this structure continue to be the object of a great many
theoretical articles, whose results, incidentally, do not always agree with the (difficnlt) measurements
that can be made.
3.1 The Static Properties of a Granular Pile 55

Randomness of the Contacts. Real-life granular materials are never made of


particles that are identical to within a few microns. But we know that the forces
of contact between two solids exert themselves over distances of the order of a
micron (see Section 2.2.1). The stacking of idealized cannon balls results in what is
referred to as hyperstatic equilibrium, involving six points of contact for each ball,
while elementary static mechanics requires only two such points located below the
center of gravity of each ball. As a result, we can afford to randomly lose several
contacts for each ball-because of variations in shapes and diameters-without
jeopardizing the stability of the entire edifice, which might even continue to give
the appearance of an ordered structure (to within a few microns). This situation
gives rise to a problem known as the percolation of contacts and leads to an almost
random spatial distribution of contact points, even with frictionless spheres [50].
If we now introduce friction, the conditions for stability are relaxed even further,
and the distribution of contact points are even more random. In short, a stack
of cannon balls, that appears at first sight highly ordered, is in reality necessarily
disordered from the point of view of the geometrical pattern of the contacts. Given
the large number of ways to dispose of the nonessential contacts, the equilibrium
of a pile of real granulars is a priori a problem with an indeterminate solution
(because of multivalued solutions).
We may even anticipate that the dynamic properties of collisions between par-
ticles with poorly defined surfaces is also likely to suffer from the same kind
of indeterminacy. We will in fact analyze this problem in Chapter 6 (see Sec-
tion 6.3.2). Referring to Figure 31, it is crucial to know the sequence in which
the three particles collide in order to predict what will happen next. This can be a
serious problem if the geometrical details are somewhat random, and even more
so if the surface properties are undetermined.
As we are about to see, the indeterminacy concerning the friction forces is even
more fundamental than just suggested.
Randomness of the Friction Forces. In the spirit of the microscopic model out-
lined in Section 2.2.1, it is instructive to ponder the fundamental role of friction
forces. It follows immediately from the three laws of dry friction that two solid
objects in contact are inevitably subject to two reactions curtailing their relative
motion. The first, which we are quite familiar with, is normal to the tangent plane at
the point of contact. It manifests itself, for instance, in the reaction of a horizontal
support offsetting the weight of an object placed on it. Except possibly for issues

FIGURE 31. Collision of three particles. Classical dynamics requires a precise knowledge
of the order in which the collisions occur, as well as of the nature of the surfaces.
56 3. Fluidization, Decompaction, and Fragmentation

related to the plastic deformations of the microcontacts, the nature and magnitude
of this reaction force are perfectly known. The tangential reaction, on the other
hand, is quite another matter. The corresponding force, called "resistance to fric-
tion" because it opposes lateral motion, is inherently indeterminate, which shatters
any hope of describing precisely and exactly the static properties of objects in fric-
tion. While it is clear that the force R opposed to gliding cannot exceed a critical
value F/-Ls, where F is the force that keeps the objects together, it is no less clear
that, as long as the system is at rest, R has an unknown value that can be anywhere
between 0 and F/-Ls. We can think of the problem by imagining that the micro-
scopic protrusions get deformed and oppose subsequent displacements, although
we do not know precisely the details of these deformations. To complicate matters
further, the deformations in question may well take place in the plastic regime,
which implies slow changes. In other words, strictly speaking, we ought to treat
the force of resistance to gliding as a time-dependent problem. It is also apparent
that a complete description of all the forces involved between two solids depend on
the prior history of their contact, specifically on how that contact was established.
The situation is illustrated in Figure 32. Part (b) of this figure shows that the
force acting on a brick placed on a support featuring two perpendicular slanted
surfaces depends on the way the brick was placed. It might have first been set
free on the left surface and subsequently pushed against the perpendicular wall, or
the other way around. The same ambiguity exists with regard to the equilibrium
of spheres resting on each other, as shown in Figure 32(c). In all these cases, the
solution to the problem depends on the prior history of the system.
It would then seem that any attempt to describe the forces of contact in a granular
edifice is doomed from the outset. It turns out that this is not always the case, as
calculations are often possible as long as enough information is available about the
way equilibrium was reached. This is in fact quite a general property of granular
materials. We will illustrate the point with a particularly simple example which
we will revisit from a different angle in Chapter 4.

Indetermination of Solid Friction: Hysteresis


We now treat the situation illustrated in Figure 32(a). A rectangular brick is placed
on a slanted surface making an angle e with the horizontal. The various forces it is

F
(b)

(c)

FIGURE 32. The forces of friction between solids can make it impossible to determine
unambiguously the magnitude of all the forces at equilibrium.
3.1 The Static Properties of a Granular Pile 57

subjected to are perfectly known. They include its weight P, the normal reaction
S of the surface, the reaction R = -kx of the spring (where x is the deviation
of its length from equilibrium and k is its stiffness, or spring constant), and the
force of friction with the surface, i.e., the tangential reaction F. The angle e can
vary between 0 and n /2. We know the coefficient of static friction fhs, and for
simplicity we assume that the coefficient of dynamic friction fhd is zero. Our goal
is to determine the compression x of the spring as a function of the angle e.
Finding the exact solution to this problem requires a bit of caution. We start
with two extreme cases. First, when e = n /2, the solution is trivial since the force
pressing the brick against the surface vanishes. Therefore, the friction force F is
itself zero, in which case we have x (n /2) = P / k. The case of a horizontal sur-
face (e = 0) is already more complicated. The solution is actually undetermined
because the spring may have been left initially in a compressed state defined by x,
with x allowed to have any value as long as the reaction R is less than the tangential
reaction due to friction. In other words, x < Pfhs/ k. Here is a case where the
position XQ of the brick cannot be determined without some additional knowledge
about the way the configuration of the system was arrived at.
Suppose now that we start with the spring at rest (xQ = 0) and we gradually
incline the surface from 0 to n /2. We denote by e+ angles that are reached by
increasing the inclination, and e- in the opposite case. The tangential component
ofthe load due to the weight increases as P sin(e+). Since the spring is initially at
rest, this load is the only force that is offset by an equal and opposite friction force,
until the moment the brick starts moving when tan(e+) = fhs. The brick then
drops down without friction (since we assumed fhd = 0) along the surface until
the spring stops in a new position Xe+, 2
which will happen when kxe2+ = P sin(e2+).
The brick being once again at rest, the friction force is reactivated, and the surface
must be inclined to a new larger angle ei to trigger the next motion. The angle
ei is such that kx e: = P sin(ei) = P sin(ei) - fhs cos(ei). We can thus see
that the descent of the brick (analogous to the stick-slip mechanism discussed in
Section 2.2.1) is marked by successive pauses determined by

~xet = sin (en = sin (ei~l) - fhs cos (ei~l)' (3-1)

As the angle e+ approaches n /2, the differences (ei~l - en go asymptotically to


zero because of the cosine dependence of the friction term.
If we now proceed in the opposite direction from vertical to horizontal, the
previous equation becomes

(3-2)

since the friction force now opposes upward, rather than downward, displacements
of the brick. In other words, the angles at which the brick stops are different on the
return trip, even though the starting and ending points are the same. The behavior
is illustrated in Figure 33.
Some simple experiments confirm the overall validity of the model just de-
scribed, although a particular behavior is often observed in the vicinity of e = n /2.
58 3. Fluidization, Decompaction, and Fragmentation

~
c
o
~
OJ
C
o
iIi

o 10 20 30 40 50 60 70 80 90

Tilt angle 80

FIGURE 33. Mechanical hysteresis effect. The successive elongations of the spring shown
in Figure 32(a) is plotted as a function of the tilt angle. The calculation assumes a static
friction coefficient f-ts of 0.3.

The mechanism of friction must, by definition, involve a zone of microscopic as-


perities operating in the plastic regime (see Section 2.2.1), even when the brick
is almost vertical. This requires that the force of contact be sufficient to break
through the very thin layer of oily residue that, barring special precautions, invari-
ably coats the surface of most ordinary materials. This condition is rarely met.
As a result, the operating point often follows the smooth sinCe) curve-at least
for a while-as the angle is decreased from Jr /2. Stated differently, the force of
friction appears to obey a certain threshold, corresponding to a minimum value of
the pressing force.
Other than this, the hysteresis effect discussed in the context of the simplified
model considered above can be generalized in the form of a recursion relation
between adjacent rest stages of the system

where Nand T are unit vectors, respectively, parallel and perpendicular to the
displacement of the object and E = ± I depending on the direction of the displace-
ment. We know that f-ts = tanCes ), where es is known as Coulomb's angle. It is
the angle of inclination required for the first slip of the brick when the tilt of the
support is increased from horizontal.
It is easy to appreciate the subtle and ambiguous character of the forces involved
in a system at equilibrium by considering the following situation. Imagine that we
tilt the support by a given angle ei greater than tan -1 Clls) and start with the spring
in a quiescent state. What will the final position of the brick be when we place it
on its support? In light of the previous discussion, it should be obvious that the
result will depend on precisely how we conduct the experiment. In particular, do
we first place the brick on its support and then let the spring do what it will, or do
we first find the equilibrium position of the spring and then let the brick down?
3.1 The Static Properties of a Granular Pile 59

Let us examine both cases:

• The contact is made with the spring first and the support second. In this case,
the spring is first compressed until kx l is slightly larger than P sinCe). The
brick is then placed in contact with its support. The spring cannot stretch
back because of the force of friction directed downward. The equilibrium
position is then given by kx l = P sinCe) +ftsP cos(e).
• The contact is made with the support first. The brick is then left to slide
down, compressing the spring. In this case, the friction forces are devel-
oped first. They oppose motion toward the bottom, and are therefore di-
rected upward, opposite to the weight. The spring will contract until kX2 =
P sinCe) - ftsP cos(e).

The two equilibrium positions are obviously not the same. It is impossible to
predict the final position and the applicable forces short of knowing the details of
the sequence of steps.
There is an analogy between the behavior of this system and that observed in
magnetic compounds, in which the hysteresis is due at least in part to the irre-
versibility of the formation and disappearance of domains. A tally of the forces at
play in a granular material, even when at rest, suffers from a number of indeter-
minacies, and we can expect serious difficulties in trying to model the equilibrium
of such a structure.

Distribution of Stresses in a Granular Medium


The disorder of the contacts and friction forces inherent to any real granular stack
suggests that the behavior of such materials may not conform to what we might
expect from the point of view of its elastic properties. In particular, an external
stress applied to such a structure will follow an erratic path depending on where
particles make contact. Furthermore, as the stress is increased and the particles start
deforming, new contacts are created, thereby creating additional stress paths. For
that reason, the rigidity of the structure goes up with stress, leading to substantial
deviations from the linear behavior implied by Hooke's law. These characteristics
specific to granular materials confer on them some rather unusual properties, some
of which are put to good use in everyday applications. 2 The proliferation of stress
channels under the influence of an increasing load has been demonstrated by Dantu
in an elegant experiment illustrated in Figure 34 [51].
Dantu used a very idealized granular medium made of glass cylinders placed
in a transparent container with a flat bottom and vertical walls. He stacked these
cylinders with their axes all lined up in the same direction, perpendicular to the
plane of Figure 34. Upon compressing the stack with a piston, he was able to
observe the stress pattern by lighting up the entire assembly from behind and
using crossed polarizers placed on either side of the container. This method,

2For instance, gravel ballast used to stabilize railways has very nonlinear elastic properties. Because
it becomes stiffer when compressed, it can offer a range of malleability and strength for a wide range
of loads.
60 3. Fluidization, Decompaction, and Fragmentation

FIGURE 34. Stress pattern observed in a two-dimensional granular material under com-
pression (after Dantu [51]).

exploiting the photoelastic effect, is widely used in the industry. It generates a


kind of map of the stress patterns in the material as it is subjected to various
loads.
As can be seen in Figure 34, the resulting image contains a wealth of information.
Besides verifying that an increase in the load causes more or less random changes
in the intricate network of stresses, with an overall increase in their density, it
also demonstrates vividly that the stress tends to be redirected laterally toward the
vertical walls. This behavior is markedly different from that of a homogenous
solid under vertical compression. Indeed, the fact that the stress field tends to be
redirected perpendicularly to the initial load is a defining characteristic of granular
materials. Even when gravity is the only acting force, this very property gives rise
to important and spectacular phenomena known as arch or vault effects, which we
have already mentioned on several occasions.
With the help of a particularly simple example shown in Figure 35, we begin
by deriving the stability conditions of an arch made of granules subject to friction,
when they are piled on top of each other and anchored at the bottom on two
horizontal supports. Our goal is to determine what arrangement of such particles
gives maximum stability to the structure. We will distinguish two cases. In the
first, particles are in equilibrium under their own weight. In the second, a load is
distributed over the entire edifice.

Arch in Equilibrium Under its Own Weight


Let p be the linear density of the chain, and let F A and F B be the two forces acting
on the extremities of an elemental segment dl, as shown in Figure 35. The figure
helps us understand that the arch will be stable if no torque or shearing force acts
3.1 The Static Properties of a Granular Pile 61

i~
F.
J1~dl
a p gdl

FIGURE 35. Contact chains involved in the formation of an arch.

on any of the contact points between contiguous spheres. Some shearing force is
actually tolerable, as long as we remain in the dry friction regime, which requires
particular conditions that are left as an exercise. For simplicity, we simply seek
here the conditions that make these forces, as well as the torque, null. We do this
by writing that the resultant of all the forces involved must be tangent to the curve
at points A and B. Assuming that the segment dl is rigid, we may write

FA +FB + pgdl = O.

e
If is the angle the segment dl makes with the horizontal, the following relations
hold in a Cartesian coordinate system: dxjdl = cos(e), dyjdl = sinCe), and
dx 2 + d y 2 = d1 2 . The equilibrium equation projected on the horizontal axis yields

dx
F - =Fh (3-3)
dl '

where Fh is the horizontal component of the force of tension experienced by the


arch. This constraint is imposed by external conditions. We also have in projection
against the vertical axis

(FdYdl ) A
_ (F dYdl ) B
_ pg dl = 0,

which, by differentiation, gives

Y
-d ( Fd- ) +pg=O.
dl dl

Combining this last equation with (3-3) leads to

d(d-dxY) +pgFh- -- 0 .
-
dl
(3-4)
62 3. Fluidization, Decompaction, and Fragmentation

We thus arrive at a differential equation describing the curve we are looking for.
It reads

pg dt
---
Fh dx
whose well-known solution is

y = - ;; [COSh (- ;~ x ) - I} (3-5)

As we might have guessed, this last result is similar to the equation describing a
flexible rope hanging loose between its two suspended ends. The requirement of
"flexibility" means in the present context that the particles the arch is made of may
roll on each other but are not allowed to slip laterally. In other words, (3-5) correctly
describes a granular arch if-and only if-the constituent particles are subject to
rolling without gliding, which implies some restrictions specified in Section 2.2.2.
Figure 36 shows a magnificent example of architecture that owes its stability to the
principles we have just discussed. Essentially the same approach can be followed to
derive the equation corresponding to an arch supporting a load whose distribution
is known. We will treat the simple case of an evenly distributed weight.
Arch Supporting an Evenly Distributed Load
We now assume that the arch has a negligible linear density, but that it supports
an evenly distributed linear mass mo. In this case, we have mag dx = pg dt, and
(3-4) must be modified to

This leads to the second-order differential equation


2
d y mag _ 0
dx 2 + F -
h
,

whose solution is obviously an inverted parabola. The highest point of the parabola
has a horizontal tangent, from which we get
1 mag 2
y = -z--x
Fh
The force of tension at a point of abscissa x is straightforward to calculate. The
result is

A similar derivation in three dimensions leads to a dome-shaped structure whose


profile follows a y ex x 3 dependence. A recent simulation based on somewhat
different arguments came up with a similar result [52].
3.1 The Static Properties of a Granular Pile 63

FIGURE 36. Photograph ofthe aqueduct ofMaintenon, built during the reign of Louis XlV.
It is a magnificent example of arches that have remained stable through several centuries.
64 3. Fluidization, Decompaction, and Fragmentation

3.1.2. Stress-Strain Relations


Identical Spherical Granules
It is useful to recall here a result concerning a theoretical lattice with perfectly
ordered contact points and friction, exemplified by a compact triangular array in
two dimensions or "canon ball stack" in three dimensions. We have seen that such
an arrangement obeys Hertz's law of penetration (see Section 2.2.2), which states
that the relative deformation d of two spheres in contact subjected to a force f
goes as f ex d 3 / 2 .
Generalizing, the relative deformation D of a perfect lattice of spheres under
the influence of a vertical load F is such that F ex D 3 / 2 . This is to be contrasted to
Hooke's law, valid for an elastic solid, which, of course, states that F ex D. Even
without the complications associated with the disordered pattern of the contacts
and of the distribution of friction forces, the stress-strain relation of a stack of
spheres is already nonlinear.
As we have mentioned previously, the proliferation of contacts under the effect
of an applied stress can only exacerbate the nonlinearity ofthe response. A classic
article on this topic can be found in the literature [42]. Furthermore, the randomness
and-more importantly-the indeterminacy of the stress pattern in the presence
of friction constitute difficult obstacles to overcome. That is why we shall neglect
friction entirely in the following discussion. We will retain only the disorder
associated with variations in the diameter of frictionless granules.

Granules of Different Sizes: Power Law and Electrical Analogy


Consider a granular pile made of particles whose radii range from r - dr to r + dr
(as we are likely to deal with in real life), the average value r being itself relatively
small. 3 Although such a pile can give the appearance of a "perfect crystal," we
have already pointed out the importance of the quasi-randomness of the contact
paths and of the network of stresses.
We now focus our attention on two particular granules (or canon balls) that
are in close proximity to each other but not physically in contact. Increasing the
load applied to the overall structure will cause the two granules to get gradually
closer, until contact is finally established. At that moment, a new stress path is
created. Decreasing the load, on the other hand, can never result in new contacts.
These comments suggest an electrical analogy with the current-voltage transfer
characteristic of a diode, as illustrated in Figure 37.
The analogy has undeniable virtues from a pedagogical perspective. It suggests
a methodology enabling us to study the problem and establishes a connection
with problems whose theoretical and practical solutions have long been known.
It should not, however, be pushed too far, as the nature of the quantities involved
is different. Currents and voltages are vectors, while the deformation D and the
external stress F are usually second-rank tensors. This turns out not to be unduly
restrictive for the particular configuration we have chosen to treat. We can then

3A more detailed discussion of this problem can be found in an article by S. Roux [24].
3.1 The Static Properties of a Granular Pile 65

c-
o

E
.2
OJ
8

v
(External Force)

FIGURE 37. Transfer function of an electrical diode. It is analogous to the stress-strain


characteristic of a network of particles with different sizes.

describe the behavior of the system in terms of 1-V characteristics

1= c(V - Vt), when V> Vt,


1=0, when V:S VI,
where the constant c is the "conductance" when contact is established, and Vt is a
threshold voltage.
The buildup of contact paths under increasing stress can be thought of in terms
of the following electrical analogy. Consider a two-dimensional network of diodes
whose threshold voltages Vt are randomly distributed around some median value.
As we apply an increasing voltage across the entire network, more and more
diodes are turned on, and new current-carrying channels percolate from a given
diode to the next. A theoretical analysis of this problem is quite difficult because of
the combination of two random variables-the distribution of threshold voltages,
and the spatial distribution of the diodes. Nevertheless, computer simulations
have been carried out [53]. They produced results in fairly good agreement with
experimental results [54]. The observed behavior is

F ex: (D - D o)3.S.

This dependence is decidedly nonlinear. In addition to the existence of a threshold,


the exponent is ~-substantially larger than in Hertz's law, which has an exponent
of ~. We might add that experiments have failed to confirm the existence of a
Hertz regime (with a ~ power law) expected theoretically when the stress is high
enough for all the particles to have come into contact. In any event, the material
does become more rigid as it is being compressed, and the present model suggests
that this effect is related to the proliferation of contact points.

3.1.3. Second Principle: Reynolds's Dilatancy


In an article published in 1885, Reynolds observed that "a strongly compacted
granular material placed in a flexible envelope invariably sees its volume increase
66 3. Fluidization, Decompaction, and Fragmentation

Tube filled with


colored liquid ----.. !

Rubber pouch
containing sand
+ colored liquid

FIGURE 38. A simple demonstration of Reynolds's dilatancy principle. A rubber balloon


is filled with coarse sand and a colored liquid. A thin glass tube, itself half filled with the
colored liquid, penetrates the balloon down into the sand. Contrary to intuition, the level
of the liquid in the tube drops when the balloon is squeezed. The explanation is that the
initially compacted sand expands when compressed.

when the envelope is deformed. If the envelope is unstretchable but deformable,


no deformation is possible until the applied force breaks the envelope or fractures
the granular material" [17]. This observation has remained until now one of the
great principles of the physics of granular materials. It has come to be known as
Reynolds's dilatancy principle. It readily manifests itself in some simple experi-
ments which we proceed to describe.
As we walk on a wet sand beach, most of us may have noticed that the sand
appears to dry up around our footprints. This phenomenon can be explained on
the basis of the dilatancy principle: As our feet transmit our weight to the ground,
the sand reacts by increasing its volume locally and soaking up the surface water,
giving the illusion that it dries up. Figure 38 shows another simple experiment that
can be done with a balloon filled with sand and a colored liquid. It too seems to fly
in the face of intuition, until we realize that the sand expands when compressed.
It is important to emphasize that Reynolds's principle must be adhered to in its
entirety. In particular, the stipulation that the material be "strongly compacted"
is absolutely essential. The purpose of the following discussion is to show, with
the help of an especially simple and perhaps somewhat reductive example, the
limitations of the dilatancy principle as it was originally stated by Reynolds.

Deformation of a Simple Parallelogram


We proceed to calculate the deformation of an extremely idealized granular, namely
four disks arranged as shown in Figure 39. The only restriction imposed by our
model is that the four disks remain in contact at all times during the deformation.
In real life, such arrangements are encountered in two-dimensional stacks of small
beads with slightly different sizes.
3.1 The Static Properties of a Granular Pile 67

FIGURE 39. Parallelogram used to model the unit cell of a two-dimensional uniform
granular medium.

The lines connecting the centers of the four disks form a parallelogram that
changes shape when forces are applied in the manner shown in Figure 39. We
focus our attention on the closed geometrical figure composed of the four disks
themselves (assumed to be rigid) and the void between them. More particularly,
we are interested in how the surface area SI of this figure varies during the defor-
mation. Let h v and hi be the lengths of the two diagonals of the parallelogram.
We can show that the surface area of the four disk sectors situated within the paral-
lelogram is constant and equal to rr R 2 . It follows that SI has the following simple
expression

(3-6)

The diagonals hi and h v are restricted to values such that ht + h~ = l6R 2 . The
variation /::"SI of the surface area of interest can then be written in terms of one of
the two diagonals only, say hi

where hi can take on values only between 2R (when the left and right disks
touch each other) and 4R cos(rr /6) (when the top and bottom disks come in con-
tact).
Figure 40 is a graph of /::"SI (normalized to 4R 2) plotted against the length hi
of the horizontal diagonal (normalized to 2R). The graph differs markedly from
what would be expected on the basis of the mechanical properties of homogeneous
and isotropic solids. As the horizontal deformation begins in response to a vertical
stress, the curve /::"St(h l ) starts out by increasing, until it reaches a maximum.
In this regime, the volume of the material goes up. This behavior agrees with
Reynolds's principle, and runs counter to what is observed with the usual homoge-
neous solids. We should point out, however, that the material does not necessarily
comply with the requirement that it be "strongly compacted," except to the extent
68 3. Fluidization, Decompaction, and Fragmentation

OJ
a: 1.0
""
~ ""
ell
"" Solid

"
~ regime
Reynolds's " "I
ell 1
CD 0.9
()

{g regime
:::J I" "
CfJ
I ""
I
""
0.8
1.0 1.2 1.4 1.6 1.8
Distance between particles (h e/2R)

FIGURE 40. Total surface area of the object defined in the text (normalized to 4R 2 ) plotted
against the horizontal distance between spheres (normalized to 2R). The inclined straight
line describes the behavior of a traditional homogeneous two-dimensional solid, whose
surface area always decreases when compressed.

that adjacent particles are already in contact. To the right of the maximum, on
the other hand, the volume does decrease, in accordance with the classical laws of
elasticity. The conclusion is that Reynolds's principle can, indeed must, depend
in a subtle way on the mode of stacking of the particles. We begin to appreciate
how ambiguous the statement "strongly compacted" really is. Finally, we have
ignored what might happen toward the left side of Figure 40. We can at least con-
ceive that a "loose" granular medium should see its volume initially shrink as it is
being compressed. Figure 41 illustrates that some arrangements obey Reynolds's
principle, while others do not.
We may attempt to calculate effective deformation parameters by resorting to
the standard methodology used in the mechanics of solids. Since we have no idea
what the elastic stiffness of the object considered might be, except that it must be

(a) (b)

16x16 16x16
Planar square lattice Triangular lattice

FIGURE 41. Two possible lattice configurations in two dimensions. Stack (a) has minimum
compactness; its volume decreases upon application of any external stress. By contrast,
stack (b) exhibits maximum compactness; it does conform to Reynolds's dilatancy principle.
3.1 The Static Properties of a Granular Pile 69

anisotropic, it is out of the question to try to define effective elastic constants. What
we can do, however, is to define the equivalent of Poisson's coefficient (]". In the
case of a homogeneous and isotropic material subjected to a vertical load, Poisson's
coefficient is defined as the ratio of the horizontal dilatation strain Uz = dhz/ h z to
the vertical contraction strain Uv = dh v / h v
Uz dhz h v
(]"=-- = - - - - .
Uv dh v hi
In the case of the parallelogram considered here, the restriction hi + h~ = 16R 2
implies that hi dhz + h v dh v = 0, which leads to an effective Poisson coefficient
given by

When the two disks on the horizontal axis are in contact, we deal with a compact
triangular stack in two dimensions (which is equivalent to a compact hexagonal
stack in three dimensions). In this case, the effective Poisson coefficient is equal
to 3. This value is abnormally large when compared to the usual solids for which
thermodynamic stability considerations impose the restriction (]" :s 0.5. 4 The co-
efficient (]" decreases as the two disks lined up vertically get closer to each other.
When the parallelogram becomes a square, which conesponds to the maximum of
the curve in Figure 40, we have (]" = 1. This can be demonstrated by differentiating
(3-6), which gives

dS! = !(h zdh v + h v dh z).


Combining this with the equations defining (]" leads to

dS! = !h zdhvCl - (]").

It is clear from this last equation that the variation in total surface area d S! changes
sign when (]" = 1. In other words, Reynolds's dilatancy principle ceases to be valid
as soon as the effective Poisson coefficient becomes smaller than 1.

Deformation of a Row of Parallelograms Placed Between Two Walls


Containers playa crucial role in the physics of granular materials, as we will ob-
serve many times in the remainder of this chapter. We might point out, incidentally,
that, although we did not spell it out, the model discussed above implicitly assumes
the presence of vertical walls to prevent particles from squirting out sideways. It
also requires dry friction to ensure the stability of the overall structure. 5 The walls

4This does not imply that the principle of thermodynamic stability does not apply here. The large value
of the coefficient (Y simply results from the local anisotropy of the material considered.
5 As an exercise, the reader can work out the stability conditions of the structure by introducing a
coefficient of static friction between adjacent disks. It turns out that the structure will collapse when
the force applied on the topmost disk exceeds a critical value determined by Coulomb's angle of
friction.
70 3. Fluidization, Decompaction, and Fragmentation

must, of course, be deformable. They effectively play the role of Young's mod-
ulus by resisting lateral movements. In the interests of simplicity, and without
undermining the argument, we assume that the vertical walls undergo a uniform
horizontal displacement Uz = (5z E, where (5z is the stress exerted by the disks on
the walls and E is Young's modulus of the material the walls are made of.
Stability having been ensured and dry friction neglected (the disks are assumed
perfectly rigid and smooth), we are ready to focus on the problem we want to solve:
What is the stress exerted on the lateral walls of the container when a vertical stress
(5 v is applied to the system?6 For the time being, we consider a system made of only
three layers of a compact triangular stack of the type illustrated in Figure 41 (b). The
reason we choose to do so is that such a stack is made of a series of parallelograms
similar to the ones we have dealt with in our previous example. We will see later
how to solve the problem for a large number oflayers (see Section 3.1.4).
By itself, an individual parallelogram exerts no resistance to the stress applied to
the disks lined up vertically. 7 It merely redirects the displacements (or strain in the
language of continuous mechanics) toward the lateral walls. Since the system is
nondissipative, the principle of conservation of work applies and the stress exerted
on the lateral walls is immediately seen to be such that

(3-7)

In light of the definition of the material's Poisson coefficient (5 given earlier, we


have
1
(5z =- .jCi(5v = K(5v. (3-8)

We have just introduced a coefficient K, which we call the coefficient ofredirection


toward the wall of a vertical stress applied to the material. s Note that K is inversely
proportional to the square root of the effective Poisson coefficient. In the case of a
compact triangular stack, its value is approximately 0.58. As we shall see shortly,
the coefficient K reflects the meshed character of a pile and plays a fundamental
role in the physics of granular materials.

3.1.4. Cylindrical Container: Janssen's Model


Generic Model: The Silo Problem
Janssen proposed a heuristic model based on the "mechanics of continuous media"
as early as 1895 [15]. His starting point was the observation that granular media

6There should be no confusion between Poisson's coefficient, which is a scalar, and stress, which is a
tensor.
7 Strictly
speaking, we ought to speak in terms of forces rather than stresses. But we have deliberately
opted to resort somewhat loosely to the traditional terminology in the mechanics of continuous media.
8We will often use the word "redirection" in this book. It means that part of the vertical stress applied
to a granular material creates a horizontal component that presses against the vertical walls. The word
"reorientation" is also sometimes used.
3.1 The Static Properties of a Granular Pile 71

had a marked proclivity to redirect vertically applied stresses toward the sides, as
described above in the context of simple models. 9 The model was elaborated on a
few years later by Lord Rayleigh [16]. It rests on two principles that are important
to understand clearly:

• From a mathematical vantage point, the medium is treated as though it were


continuous. This approximation is highly arguable in a granular medium.
Nevertheless, it makes it possible to write differential equations whose solu-
tions describe the behavior of these materials with rather surprising
accuracy.
• A vertical force Pv (or stress) applied to the material automatically generates
a proportional horizontal force Ph (or stress) such that Ph = K Pv.

This is entirely consistent with our previous analysis of a string of parallelograms.


Janssen and Rayleigh postulated that the behavior we found earlier could be ex-
tended to more complex structures than those we have considered so far. To
treat a granular medium as if it were continuous may strike some as too crude an
approximation. A few comments are in order.

• Since the material is made of discrete constituents with a finite size, the
relevant space variables by definition cannot approach zero, and writing dif-
ferential equations obviously raises serious questions. However, this type of
approximation has been used successfully in other branches of physics and
has led to solutions that for the most part seem to agree well with experiments.
A case in point is the transition from the mechanics of atomic structures to
that of continuous media. It should be kept in mind, however, that the ap-
proach works only as long as we deal with large numbers of particles. It
breaks down when we try to explore properties on a local level, such as when
the number of particles involved is small. We will see an example of this
limitation when we study the dynamic behavior of granulars, where a model
based on continuous media fails to account for the phenomenon of fragmen-
tation in granular media that are vibrated or in forced flow. The message
here is that the approach applies only to structures made of a large number of
partic1es. lO
• Problems related to the lack of cohesion of a dry granular medium are even
more fundamental. By definition, a granular material is heterogeneous and
contains empty spaces. Given the constraints we have specified early on in
this book, namely that the particles not interact with their gas environment, the
intergranular voids inevitably impose some restrictions on how deformations

9This property is not the exclusivity of granular materials. Soft media, such as lUbber, exhibit it
too. Unlike lUbber, however, the granulars we will deal with have no cohesion. We will see some
consequences of this fact a little later in this chapter.
10 Apractical method useful in simulation work will be presented in Section 6.1.3. It permits a
smooth transition from discrete variables (position, individual velocity, mass) to continuous variables
consistent with thermodynamic quantities (density, mean speed, temperature, etc.).
72 3. Fluidization, Decompaction, and Fragmentation

and stresses can propagate within the medium. A differential quantity such as
a variation in pressure becomes meaningless the moment the particles are not
in contact. Likewise, the concept of individual rotations presents difficulties.
Classical mechanics, as it applies to continuous and homogeneous solids, can
handle shear forces routinely but is powerless to deal with local rotations. We
may anticipate that a description based on the theory of continuous media
requires that the network of forces be transmitted throughout the medium
without interruption. A variable such as pressure can be defined in a granular
material only inasmuch as the contact chain is unbroken and rotations of
individual particles do not occur. We will make use on these remarks in what
follows, notably in our analysis of the fragmentation of granulars in forced
flow (see Section 3.2.4).
• Problems related to the finite size and discontinuities of these materials can be
particularly bothersome when it comes to their dynamic properties, as we will
appreciate later on. Given the wide range of relaxation times in such systems,
a continuous theory may be useful for sufficiently long-term observations,
giving the system plenty of time to relax, but completely inadequate for
short-term observations. Indeed, it is not all that unusual in the world of
physics for phenomena and their description to be quite different depending
on the duration of the observations. We will see examples of this in this
chapter.

Referring to Figure 42, we consider a sheet of thickness dh situated at a height h


in a cylinder of surface area A and perimeter P. Such a sheet is in equilibrium
under the combined effect of several forces.

• Since pressure increases with depth (we adopt the convention that h = 0 at
the top of the cylinder, and h > 0 toward the bottom), the slice of interest
experiences a force directed toward the top and equal to A dpv'
• The weight of a slice of thickness dh constitutes a force directed toward
the bottom and equal to pg dAh, where p is the volumetric density of the
material, assumed constant throughout the slice.

Surface Area A

Perimeter P

FIGURE 42. Cylindrical configuration defining the parameters used in Janssen's model.
3.1 The Static Properties of a Granular Pile 73

• The forces of friction with the walls, resulting from an infinitesimal move-
ment of the slice toward the bottom, are directed upward. This is not an
arbitrary choice, as it amounts to assuming that the material slowly settles
under the action of gravity, a trend opposed by friction. 11 As we will see later,
there are some experimental situations that would justify making the opposite
choice for the direction of the forces of friction. In any case, the relevant force
exerts itself all around the wall, over an area P dh. Its value is thus equal
to fLs Ph P dh. Taking into account (3-8), which relates the horizontal and
vertical components of the stress, the forces of friction become KfLs Pv P dh.

We are now in a position to write the equilibrium condition for the particular slice
considered. It reads

Adpv + KfLsPpv dh = pgA dh.

Dividing through by dh, we obtain the differential equation

(3-9)

It can be rewritten as

which can be readily integrated to

(3-10)

where C is a constant to be determined from the initial conditions. Assume, for


instance, that we apply a pressure PvO on the top of the structure by placing there
a mass M distributed over the surface area A, we have PvO = M g I A. Under these
conditions, the constant C is equal to (Pvo - pg AI P KfLs), and (3-10) becomes

Pv = pg P:fLs [1 - exp (-KfLS : h)] + pvoexp (-KfLs~h). (3-11)

Equation (3-11) is a slightly generalized form of Janssen's equation.


It is instructive to examine the consequences of (3-11) when PvO = O. The trend
is illustrated in Figure 43. For small values of h, i.e., near the top of the structure,
the pressure goes as Pv :::::0 pgh, which corresponds to the conventional hydrostatic
pressure, similar to that exerted by a column of water of depth h.

llThis is a crucial point to understand. Janssen and Rayleigh implicitly considered that the forces
of friction were always right on the edge of giving way at the walls and throughout the structure.
This picture completely ignores the indeterminacies affecting the forces of friction which we have
discussed in Section 3.1.1. In other words, both authors replaced the inequality T :: /LsN by
the equality T = /LsN-a rather bold assumption. Another way to put it is that the calculation
corresponds to a fully relaxed pile. As such, the model applies to long-term observations. We shall
see later that reality can be considerably more complex, particular for short-term observations.
74 3. Fluidization, Decompaction, and Fragmentation

Vertical Pressure
o ...-- Hydrostatic Psat
regime

1:
0>
'iii
I

Saturated
regime
h

FIGURE 43. Dependence of the vertical pressure in a silo as a function of height.

When h becomes larger than about A/ P Kf.-L" the vertical pressure Pv saturates,
as it approaches asymptotically a limit given by Pv ---+ pg(A/ P Kf.-Ls).
Denoting by X the argument appearing in the decreasing exponential in (3-11),
we have X = (Ph/ A)Kf.-Ls, where Ph represents the outer vertical surface area of
the container, while A is its cross-sectional area. Because it will reappear later on,
we define the ratio of these two areas as a parameter S, which we call the aspect
ratio (S = Ph/A). In terms of this new parameter S, the argument, which we
denote X, appearing in the exponential is simply X = S Kf.-Ls. For reasons that will
become clear later on, we call this argument X the decompaction parameter. It
is a dimensionless number that completely characterizes the distribution of forces
in a cylindrical pile. To show that this is true, consider the mass m of granular
material contained within a depth h counted from the top; it is given by m = pAh.
The vertical force F v exerted on a layer at that particular depth h is equal to
_ _ mg -x
Fv - Pv A - -(1 - e ). (3-12)
X
Through the combined effects of dovetailing and friction with the walls, the ap-
parent weight of the cylindrical column is reduced by a factor that depends solely
on the dimensionless decompaction parameter X.

Specific Applications
We briefly consider some specific geometries shown in Figure 44.
Cylindrical Container ofDiameter D
Equation (3-11) then becomes

Pv(h) = p gD- - [ 1- exp ( - 4Kf.-Ls


- - h )] .
4Kf.-Ls D
3.2 Dynamic Properties of a Granular Pile 75

2-D Cell 3-D Cell

FIGURE 44. Two- and three-dimensional container geometries.

Two-Dimensional Container
A single layer of a granular is confined between two large flat frontal plates. The
granular medium has no friction with those plates, although it does with the small
end walls. We will make repeated use of such a two-dimensional stack in the
remainder of this book. Let e be the thickness of the material, L the length of the
cell, and h its height (see Figure 44). The aspect ratio S of such a stack is given
by

Ph 2he 2h
S------ (3-13)
- A - Le - L'

The numerator includes the active part of the perimeter only, ignoring that which
does not contribute to friction. The decompaction parameter reads X = SK!J,s =
2K!J,shjL, and (3-11) becomes

Pv(h) = p gL- - [ 1 - exp ( - 2K!J,s


- - h )] .
2K!J,s L

We will invoke this result in Section 3.2.3.

3.2. Dynamic Properties of a Granular Pile


As we have no doubt come to appreciate from the simplified analysis just con-
cluded, the static properties of something as innocuous as a sand pile can be quite
complicated. The disorder of the contact points combines with the randomness
and indeterminacy of the friction forces to give rise to peculiar nonlinear properties
that are characteristic of granular materials. We now tum our attention to the dy-
namic behavior of these materials. It does not take much prescience to anticipate
that phenomena such as arch formation, nonlinear elasticity, hysterisis effects, to
name but a few, are likely to confer on granulars some rather unusual properties
here as well. We begin by examining the elementary problem of fluidization and
76 3. Fluidization, Decompaction, and Fragmentation

(a) (b)

N
spheres

FIGURE 45. Spheres excited by a vibrating plate. Diagram (a) is the classic problem of a
"bouncing ball." Diagram (b) corresponds to several spheres stacked vertically.

decompaction of these materials when they are subjected to vibrations. 12 These


stimuli turn out to be particularly simple to model theoretically and study experi-
mentally. Spontaneous avalanche flows will be deferred until Chapter 4.
The transition from static to dynamic properties brings into the picture a new
interaction which we considered briefly in Section 2.2.2. We are referring to col-
lisions, which can playa predominant role, possibly even more important than
friction, particularly if the particles involved have a high coefficient of elastic
restitution. Accordingly, we proceed down our logical progression by examining
the simplest possible case, namely that of a column of spherical particles stacked
on top of each other as shown in Figure 45. We assume that they are devoid of
friction with the walls and are subjected to a sinusoidal vertical vibration. The
analysis to follow, simplified in that it considers only collisions to the exclusion of
friction, will enable us to extract a few general principles which will prove useful
in treating the cases of two- and three-dimensional stacks.

3.2.1. A Column of Spheres Subjected to a Vertical Vibration


Analyzing the behavior of a column of spheres is of interest for two reasons [37],
[55], [56]. First, the exercise reveals a serious difficulty in modeling collisions
when the particles involved coalesce into a cluster. Second, it shows the existence
of two distinct regimes-one in which the particles remain stuck together and
move collectively as a solid bloc, as it were, and the other in which they encounter
each other only sporadically during brief collisions and spend most of their time
spread apart.
Before deriving equations and working out their solutions, it useful to keep in
mind some orders of magnitude.

12The importance of making a distinction between these two terms will become apparent as we go
further in our discussion. Fluidization of a dry granular gives it dynamic properties reminiscent of
those of a nonviscous liquid or gas. Decompaction is a phenomenon that gives it the ability to execute
internal movements of reorganization, for instance by convection.
3.2 Dynamic Properties of a Granular Pile 77

Some Orders of Magnitude


As we have seen in Section 2.1, the desire to limit the relevant interactions to
collisions and friction restricts us to particles at least 100 p,m in size. Consider
commercially available beads, with a typical diameter of3 mm. This size defines a
z
length that is characteristic of the problem. For obvious reasons of convenience,
it is desirable for the trajectories of the beads to be of the same order of magnitude.
Since the problem is governed by the acceleration g of gravity, fixing a spatial scale
amounts of fixing the time scale of ballistic flights, which is given by tv = 2.J2z / g
and is equal to a fraction of one-tenth of a second. The vibrating plate must operate
consistently within that time scale, which means that its frequency must be a few
tens of Hertz. Its motion is described by A sin(wt). The amplitude A must be
suffcient to propel the beads up in the air. That implies that the acceleration that
the plate imparts to the beads must exceed that of gravity. We have pointed out
previously that it is convenient to introduce a normalized acceleration r related
to the real acceleration y by r = y / g. In the present case, the real acceleration
is equal to Awl, and the normalized acceleration is r = Awl/ g. It all boils down
to choosing some practical parameters. Typically, we might use a frequency f =
w/2n of 20 Hz, in which case A = 0.64 mm, and the maximum velocity of the
plate is Aw = 8 cm/s.

MathematicalAnalysis of the Problem


The simplest way to model the problem is to solve, on an event-per-event basis,
Newton's equations [55], [56]. We can work out an iterative description involving
equations of ballistic movements, in conjunction with the equations governing the
collisions themselves [(2-2) for collisions between beads, and (2-3) forthe collision
between the first bead and the vibrating plate]. The ballistic equations read

(i = 1,2,3, ... , N), (3-14)

where ZiQ and Via are measured immediately following the previous collision. The
collisions are described by linear equations of the form

(3-15)

where Ui and Vi are the velocities of bead i immediately before and immediately
after a collision measured in the frame of reference attached to the center of mass
of the two colliding objects. By convention, Ua and Va refer to the vibrating plate.
The beads are numbered sequentially from bottom to top.
How to implement such a calculation constitutes an important physics problem.
It involves making some decisions on the sequence of events-in this case ballis-
tic flights and collisions between the N particles and the vibrating plate. Without
going into the details of the computer simulation, which will be covered in Chap-
ter 6, we will simply mention the simplest and most natural algorithm. It belongs in
the class of so-called "event-driven" algorithms, which analyze events sequentially
78 3. Fluidization, Decompaction, and Fragmentation

10 .-.:::,....----.----y---,...---,-----,

N
o -",
.........
..........

0.00 0.02 0.04


Time (s)

FIGURE 46. Dynamics of a collection often spheres, calculated with the help of (3-14) and
(3-15), using 8 p = 0.6 and 8 = 1. The upper diagram shows the positions of the spheres
plotted against time. The lower diagram gives the time intervals between successive colli-
sions (after [59]).

in the order in which they occur. 13 For instance, starting at some time t, we can
construct a matrix T which contains a sequence of times ti at which future events
are predicted to take place. In particular, the smallest element tim of that matrix is
the time of the very next event. When it has occurred and its consequences have
been calculated, a new matrix T is constructed, and the cycle is started all over
again. There is absolutely nothing wrong with this procedure as long as we are
not faced with two or more simultaneous events (in the sense of physics or of the
computer's ability to distinguish them), in which case some inherently arbitrary
choices have to be made. Erroneous decisions can, of course, seriously taint the
subsequent chain of events. More pragmatically, the reader may recall a calcula-
tion we did in Section 2.2.2, which convinced us that collisions take some finite
amount of time (of the order of a microsecond in the example treated), and that
this time further depends on the velocity of the colliding objects. Including this
effect leads to additional uncertainties.
Summarizing, the calculation technique based on Newton's equation for bal-
listic flights (3-14) and event-driven collision matrices (3-15) can run into major
problems when the sequence of events becomes difficult to keep track of. The point
is illustrated in Figure 46, which tells the story in the case of ten beads. The figure
shows the trajectories Zi (t) over the duration of one excitation period, as well as the
time intervals separating sequential events. We notice immediately-and that is a
very general result-that the time interval between consecutive collisions becomes
exceedingly small (less than 10-7 s in the present case) when the ten beads get so
tightly clustered as to be practically in contact. From a mathematical point of view,
we may even claim that contact will be achieved when the interval between events

13 "Event-driven" algorithms are to be distinguished from "time-driven" algorithms and other sequen-
tial techniques. They are also sometimes referred to under the heading of "collision method."
3.2 Dynamic Properties of a Granular Pile 79

goes to zero, although that may not be too realistic because even the most simplistic
models indicate that two colliding objects remain in contact for durations as long
as several tens of time intervals in this limiting regime. Hence a serious problem.
It is clear that the difficulty has to do with the existence of two separate character-
istic times of the system-the duration fl Tc of a collision and the interval flt m ,m-l
between two successive events. The validity of the model is assured only to the
degree that fl Tc « fltm ,m-l. Some might further object that in this troublesome
regime the distance between particles becomes comparable to the deformations or
microasperities of the surfaces, which in itself may invalidate (3-15). This compli-
cation, which is a constant source of vexation in numerical simulations, has come
to be known in the literature as inelastic collapse or inelastic catastrophe [57]-
[59].14
It is possible to circumvent these difficulties by introducing the notion of block. IS
To that end, we introduce a threshold velocity V c (see Section 6.2.1) which marks
the separation between beads considered separate from those that are effectively
stuck together. When the velocities of two particles that have just collided differs
by less than vc , we treat them as if they had merged into a single block. In practice,
the velocity V c is chosen as small as possible (for instance, 10-7 m/s) while still
compatible with the precision of the computer. We may also verify that changing
the value of V c does not substantially affect the results of the calculated dynamic
behavior of the system. The key is to find solutions that remain stable with respect
to changes in V c ' It is, of course, imperative to conserve the center of mass of
a block during its subsequent movements. Since masses do not enter explicitly
into the type of calculations discussed here, they have to be handled through the
bias of the time-dependent position z(t). Having developed a criterion for block
formation, it is also important to consider the possibility that once they are formed,
blocks of several particles may disintegrate in part or in whole. We will defer until
Section 6.2.2 a more detailed discussion of this topic. Suffice it to say here that
different approaches produce results that are in substantial agreement. Figure 47
is offered as an example of an event in which particles clustered in two separate
blocks rearrange themselves in a different pattern following a collision between
the blocks treated in accordance with the so-called Largest-Relative-Velocity-or
LRV-criterion. 16
Results: Fluidized Phase and Condensed Phase
Numerical solutions of Newton's equations, obtained with the precautions dis-
cussed above, reveal several specific and generic behaviors that seem extendible
to two- and three-dimensional configurations. We proceed next to describe these
behaviors and to emphasize their differences, although the dividing line between
them is not always clear-cut.

14Needless to say, it is a catastrophe only from the point of view of developing a model. As we might
guess, the term was coined by people doing computer simulations.
15This is a common technique designed to get around problems of accumulation in numerical simula-
tions. We will encounter it again in Chapter 6.
16The LRV criterion will be discussed in detail in Section 6.2.2.
80 3. Fluidization, Decompaction, and Fragmentation

FIGURE 47. Behavior of five spheres initially grouped in two blocks of 3 and 2, respec-
tively. The horizontal axis represents time. The spheres rearrange themselves after the
collision. The simulation was done by applying the LRV criterion (see text) (after [59]).

Fluidization and Condensation as Functions ofAcceleration


When the coefficient of elastic restitution is high, as in the case of steel balls for
which s = 0.9, the key parameter is the normalized acceleration r. As shown in
Figure 48, a colunm of ten balls submitted to a strong acceleration is virtually in
a fluid state. By contrast, when the acceleration is lowered, the column remains
clustered together in the form of a single block that moves collectively in phase
with the vibrating plate. The system is then said to be "locked" to the excitation.
This behavior turns out to be quite stable over a wide range of excitation parameters
(acceleration r and frequency f). It can be understood, at least intuitively, with
the help of the following simple argument. Consider the ballistic trajectory of
a group of spheres moving as one. Since the spheres are tightly clustered, most
of the energy they receive from the vibration plate is dissipated in multiple and
repeated collisions. Under such conditions, the coefficient of elastic restitution of
the column is quite small. I? The problem then becomes identical to that of a single
inelastic ball, as depicted in Figure 45(a). The advantage of the much simpler case
is that it has been solved completely. One of its well-known properties is to evolve
toward chaos (Feigenbaum's scheme). Without dwelling on the details, Figure 49
enables us to understand the mechanisms involved.
The figure illustrates the trajectory of an inelastic object (simulating the col-
umn we were dealing with) placed on a vibrating plate whose vertical motion
is sinusoidal with amplitude A and frequency f. The object takes off when the
acceleration imparted by the plate is such that Au/ > g. It then leaves the plate
and follows a parabolic ballistic trajectory, falling back toward the plate which in
the meantime keeps on oscillating. It is crucial to realize that, since the impact is
nearly inelastic, the object does not bounce back. Rather, it "sticks" to the oscil-
lating plate until it is once again launched up into the air. We note that inelasticity
will also tend to cause a colunm of several beads to clump together, ensuring the
stability of the system.

17 Acollection of clustered particles presents a small global coefficient of elastic restitution even if each
individual member is nearly perfectly elastic. That is why a packed bag of marbles does not bounce
back when falling on a hard floor, even though a single marble does so spectacularly.
3.2 Dynamic Properties of a Granular Pile 81

(a)

(b)

FIGURE 48. Calculated trajectories of ten spheres with 8 p = 1 and 8 = 0.9. The system
is vibrated at 10 Hertz with two different accelerations ['. In (a), r = 8.0, and the system
is "fluidized." In (b), [' = 1.7, and the system is "condensed."

The following simple calculation is informative. Let t* be the time when the
normalized acceleration r becomes equal to 1. It is given by

At that moment, the object is launched up with an initial velocity v* equal to that
of the vibrating plate. All other parameters of the movement are straightforward
to calculate. We might note that, even if the coefficient of elastic restitution is not
strictly zero, the locking on the second harmonic (see Figure 49(b)) is particularly
stable. That is because the motions of the object and of the vibrating plate then
take place in the same direction, minimizing the velocity differences and reducing

(a)

(b)

FIGURE 49. Locking of the trajectory of a column of particles on the vibration of the plate.
Locking can occur on a number of frequencies, including (a) the fundamental, and (b) the
second harmonic.
82 3. Fluidization, Decompaction, and Fragmentation

the likelihood of rebound. 18 The same situation can exist, of course, at other
frequencies besides the second harmonic, including the fundamental.
Returning to the case of a column of beads, the considerations developed above
provide a plausible picture as to why a large-amplitude excitation can lead to
fluidization, as depicted in Figure 48(a). Even if we start with a situation in which
all the beads are clustered and form effectively a single block, enough energy can be
transmitted along the entire string to cause even the topmost bead to separate from
the block and initiate its own individual ballistic trajectory. 19 What can happen to
the topmost bead can a fortiori happen to any other bead in the stack as well.
Summarizing, a column of beads or other elastic granular material vibrated
vertically exhibits two distinct regimes depending on the acceleration imparted by
the bottom plate. For small accelerations, the system is in a "condensed" state,
with the beads practically in contact with each other and moving in unison. For
large accelerations, the system is in a "fluidized" state where the beads move
about individually much like particles in a gas or a fluid. In this latter state, the
acceleration Aw 2 is no longer the pertinent variable. Instead, that role is taken
over by the kinetic energy, proportional to A 2w 2 , or, equivalently, by the square
of the mean vibration velocity. The transition between the two regimes involves
portions of the system that are fluidized and others that are still condensed. All
of this has been satisfactorily simulated numerically and verified experimentally
[56], [60].
The existence of these two regimes and the transition from one to the other also
depend on the value of the coefficient of elastic restitution as well as on the number
of particles involved. This point is examined further in the next development.
Fluidization and Condensation as Functions ofHeight
As noted above in our discussion of a column of a sufficiently large number
of beads, the collision energy may be attenuated as it propagates up the string
of particles when the interactions are not purely elastic. If so, it may become
impossible to fluidize the upper portion of the system, which will then continue to
behave as a single compact block.
Computer simulations and theoretical calculations agree that the parameter con-
trolling this effect is a reduced variable defined by X = N (l - E:). The theory indi-
cates that there exists a critical value Xc for X such that Xc = If [37]. Simulations
suggest a somewhat less well-defined critical value Xc "'::: 3 [56].
The following calculation clarifies the order of magnitude concerning the re-
duced variable X. The coefficient E: is equal to 0.6 for aluminum and 0.92 for

18Tennis aficionados might appreciate that this very principle is exploited in the dreaded "drop shot,"
whose secret consists in deadening the rebound of the ball, with potentially devastating effects on
the opponent.
19That is the principle behind a toy sold in some specialty stores under the name "Newton's Cradle." It
consists of a linear array of a few metal balls individually suspended by threads. All the balls are in
contact when at rest and form a "block" in the sense defined in the text. When the rightmost ball is
pulled apart from the block and let go, it comes crashing into the remaining block, at which point the
leftmost ball is knocked away in the opposite direction. All the while, the inner balls remain still.
3.2 Dynamic Properties of a Granular Pile 83

hardened steel. We then find that the critical number of balls is 8 for aluminum
and 39 for steel. With very few exceptions, the granular materials typically han-
dled by the food or pharmaceutical industries have much a lower coefficient E;.
Therefore, in the vast majority of practical situations, the variable X is such that
X» 3, and the materials are most often in a condensed regime.

• Case X :s 3: As noted above, there is a possible transition from a condensed


state prevailing at weak accelerations to a fluidized state when the acceleration
increases.
• Case X ~ 3: Here we have a sufficiently tall column of material in which the
energy dissipation by collisions is rather substantial. The column acts as a
single block and the duration of contacts between particles can become quite
long-at any rate much longer than the vibration period. It is important to
realize that, strictly speaking, we are not dealing with a condensed state in
quite the sense defined previously, since the number of collisions between
particles approaches zero. What we do have is a truly compacted column
whose dynamic properties approximate closely that of a single inelastic ball.
Indeed, that was confirmed by simulations and experiments. Figure 50 shows
bifurcation diagrams for a set of ten aluminum balls with increasing accel-
erations. Note the agreement between experimental results and calculations
done with a coefficient of elastic restitution of 0.6.

The bifurcation diagram in Figure 50 shows a classic tendency toward chaos. In


particular, for certain accelerations, the system "hesitates" between two possible
states. Experimentally, we generally find that the system oscillates back and forth
between two such states. That is a telltale signature of the onset of chaos.

ob .J
3

:-<
t>
t>

-<
.....
..... 2
o t>
o

g~
Eo-<

t> t>

5
r 10

FIGURE 50. Bifurcation diagram of the time of occurrence of collisions. The product f x
Teall is plotted against the normalized acceleration ['. In the present case, N = 10, f = 30
Hertz, and sp = s = 0.6. Circles are experimental points, while triangles show the results
of numerical simulations. The diagram clearly shows locking on the fundamental, as well
as on the second and third harmonics (after [56]).
84 3. Fluidization, Decompaction, and Fragmentation

To summarize what we know at this point, a column of particles excited by a


vertical vibration can remain in a condensed state for three different reasons, all of
which can be traced to an insufficient amount of energy imparted to the column.
The three causes that have been identified are:

• The effective coefficient of elastic restitution is too low. This happens when
the particles are clustered and most of the collisional energy is dissipated in
a large number of impacts.
• The collisional energy is dissipated in the column because the collision pro-
cesses are inherently dissipative or because the column is tall. Either way,
the excitation cannot make its way to the top of the column.
• The ballistic flight of the column locks onto one of the harmonics of the
excitation period, in such a way that the velocities of the column and of the
vibrating plate are nearly synchronized. In this case, collisions are virtually
eliminated and no elastic energy is transmitted.

The phenomenon of the condensation of particles is not the exclusive province of


one- or two-dimensional systems. It can be observed quite generally in collections
of particles undergoing multiple collisions [57]. The tendency toward conden-
sation can be understood on the basis of the following simple chain of events:
When two particles coalesce, the resulting block has a lower coefficient of elastic
restitution than either particle taken individually. As a result, the block is more
likely to capture a third particle during a collision, creating a larger block with a
still lower coefficient of restitution, and so forth. The probability that a particle be
captured by a preexisting block is roughly proportional to the surface area of that
block. Therefore, larger blocks are likely to grow faster than smaller ones. This
phenomenon is similar to what happens during the nucleation of liquid droplets
on a surface. 2o

3.2.2. Two-Dimensional Stack ofFrictionless Spheres


We now examine some further properties of vibrated media by concentrating on
the behavior of two-dimensional granulars, a topic we have touched on briefly
in Section 3.1.4. It is easy to realize such a system in practice, for instance, by
arranging a single layer of granules between two glass plates that are subsequently
raised vertically. The granules are allowed to move freely in the space between
the two plates. With this technique, we can routinely prepare stacks containing a
few thousand steel pellets roughly a millimeter each in a frame whose area is only
a few square centimeters.
Experiments show that the overall behavior of such stacks depends to a great
extent on the coefficient of elastic restitution s, as well as on the friction coefficients
fl,pp between pellets and fl,pw between pellets and walls. It is therefore essential
to carry out a detailed analysis in terms of the micromechanical parameters of

Z°It is the reason why water vapor condenses into separate droplets on a cold windshield, rather than
in a uniform layer. The resulting pattern is referred to as breath figures.
3.2 Dynamic Properties of a Granular Pile 85

Pz
15

z+
I
I
I_~~
X

FIGURE 51. Statistics of the trajectories of elastic and frictionless beads at the surface
of a two-dimensional pile. The photograph reveals that the lower part of the pile remains
compacted. The histograms show the distribution of velocities along the two principal axes
(after [61]).

the materials involved. Computer simulations (see Chapter 6) have gradually


incorporated all these parameters and have been able to reproduce the behaviors
observed by experimenters. We propose to examine a number of configurations.
We will start with the simplest, which is a straight extension of the one-dimensional
case treated in the previous section.
Intuitively, we may anticipate that a stack of spherical pellets devoid of friction
between themselves or with the walls should behave essentially the same way as a
one-dimensional stack. As remarked previously, the only dissipative interactions
in a one-dimensional stack are collisions between particles and between particles
and vibrating plates. The same should be true of a two-dimensional structure
made of frictionless particles. This expectation is indeed essentially confirmed by
experiments, as illustrated in Figure 5l.
For moderate accelerations, a sufficiently tall stack remains compacted [61].
This is fully consistent with our previous analysis of a one-dimensional col-
umn. The reduced variable that governs the behavior of the stack now reads
X z = Nz(l - c), where N z represents the vertical extent of the structure. Also,
as in the one-dimensional case, experiments show that large accelerations cause
the stack to disintegrate. Likewise, when X z substantially exceeds the value 3,
collision waves are almost completely attenuated as they propagate up the struc-
ture. Experiments also reveal evidence of locking on the various harmonics of the
excitation frequency.
There is, however, one new phenomenon that has no counterpart in one di-
mension. When a granular medium with sufficient lateral extension is vibrated
energetically, waves of a particular type may be excited, which are themselves sub-
ject to bifurcations and can lead to chaos. They are reminiscent of the well-known
86 3. Fluidization, Decompaction, and Fragmentation

Faraday instabilities in liquids and are currently the focus of active research. We
will briefly discuss them later in this chapter (see Section 3.2.5).
The experimental apparatus described above allows us to study the nature of
the fluidized top few layers of pellets which, as shown in Figure 51, seem to
follow complex ballistic trajectories above the compacted reservoir. This can be
done by visually recording the pellets with a CCD camera over a duration of
many thousands of excitation periods. A particularly useful technique is to use a
stroboscopic lighting system flashing pulses of light at regularly spaced instants
synchronized with multiples of the excitation period. The technique lends itself
to direct measurements of the instantaneous velocities of the fluidized particles.
Subsequent data manipulation generates the statistics of the velocities at different
times during one period. The results can then be plotted in the form of histograms.
Figure 51 includes such histograms, showing the probability distribution functions
of the horizontal and vertical components of the velocities centered about their
mean values (v x ) and (v z ). The two histograms have approximately the same
width, indicating that the medium behaves isotropically, in conformance with the
symmetry of the problem. As it turns out, this is one of the characteristics of a
fluid or gas in thermal equilibrium.

Some Comments on Scaling


A remark concerning what is often referred to as scaling laws is in order. The
issue is best summarized as follows. Laboratory experiments normally deal with
relatively small-size models, while industrial scales are usually considerably larger.
The question then is: How should an experimental result derived on the basis of
a typical linear dimension L, characteristic of the experimental apparatus, be
extrapolated to a situation involving a potentially much larger typical length aL?
Based on what we know so far, it would seem that what truly matters in the world
of granulars is not so much the dimension of the hardware as the total number of
particles involved. In other words, if energy dissipation during collisions is really
what governs the physics of granular materials, the pertinent phenomenon occurs
at the level of individual particles regardless of the size of a particular piece of
hardware. As long as collisions are the dominant factor, the scaling of the problem
should go as the dimensionless ratio L / D, D being the typical diameter of the
interacting particles.
Given what we have learned from Janssen's model, it is not obvious that the
same conclusion holds for interactions by friction. It would seem, instead, that
Janssen's model should scale as the length L regardless of the number of gran-
ules in the system. Pushing the argument to the extreme, scaling might even be
more complicated when rotations are involved. This can be understood in an ele-
mentary way by considering an interaction by strong friction in a row of N balls
subject to rolling without gliding. This gives rise to a kind of self-organization
of the angular momenta of the particles, an example of which we will encounter
in Section 3.2.4. In the present case, a left-handed rotation by an angle al of the
first ball at the start of the row triggers rotations down the chain of particles in
directions that alternate between right-handed and left-handed. The ball in nth
3.2 Dynamic Properties of a Granular Pile 87

position will tum by an angle an = (_l)n+ 1a 1. In scaling this case, it appears


that we need to know, among other things, whether the new system contains an
odd or an even number of particles, which singularly complicates the task. The
same type of reasoning can be applied to the excitation of a granular material by
ultrasonic waves that are diffused and diffracted at every contact between par-
ticles. At first blush, it seems that the problem should scale as the number of
particles, i.e., as the dimensionless ratio L / D. Unfortunately, this conclusion is
contradicted the moment we realize that the absorption of an acoustic wave de-
pends on the total length traversed, and not just the number of particles. That would
argue in favor of scaling as the dimension L. In light of these remarks, scaling of
the physics of granular materials generally calls for considerable caution. It re-
quires a detailed understanding of the relevant interactions at the most fundamental
level.

3.2.3. Two-Dimensional Stack of Spheres with Friction


The introduction of coefficients JL pp between pellets and JL pw between pellets and
walls gives rise to behaviors that are completely different from those we have
just described. For instance, friction is responsible for movements of localized or
collective convection whose dynamic properties we will cover later on. Friction
also leads to the phenomenon of gradual decompaction, which we will analyze
and quantify in a simplified two-dimensional case, although it has also been ob-
served in real three-dimensional systems [62]. We proceed to study phenomena of
convection, followed by decompaction, the two being actually closely connected.
Both depend on friction between a granular material and the walls of the container.
Throughout this section, we will be dealing with a granular medium composed
of a two-dimensional stack of poorly elastic pellets but subject to friction (e.g.,
pellets of oxidized aluminum, 1.5 mm in diameter), confined between two glass
plates in the manner described previously (see Section 3.1.4). The resulting system
is to be secured on a vibrating plate delivering a sinusoidal excitation of amplitude
a sin(wt). In the discussion which follows, we will often invoke the normalized
acceleration r, such as it was defined earlier.
Before focusing on two-dimensional media for which we shall derive a fairly
detailed description to be compared against actual experiments, we start out by
deriving a number of generic equations which will prove useful later on [63].
They apply equally well to three-dimensional cylindrical structures, somewhat in
the same spirit as for those we studied in Section 3.1.4.

Generic Model
Once again we resort to the same notation we used in our discussion of Janssen's
model, particularly when we derived (3-9), in which A represents the surface area
of the cylinder's base and P its perimeter. However, for reasons of convenience,
we adopt a different convention for the heights h. From here on, they will be
referenced to the base (h = 0) and counted positively toward the top. The geometry
is illustrated in Figure 52.
88 3. Fluidization, Decompaction, and Fragmentation

A ho
Decompacted

dm
Compacted
p
Ih Ih'
tr
FIGURE 52. Gradual decompaction model of a cylindrical stack of a granular material.

As pointed out earlier, (3-9) differs from an ordinary equation of hydrostatic


equilibrium by the addition of a correction term due to friction with the wall. The
reader may recall that this equation was derived by considering that the granular
material consists of stacked homogeneous sheets rubbing against the lateral wall,
and that the friction was assumed to be on the verge of giving way throughout.
The entire system is subjected to a vertical acceleration directed upward. A sheet
of mass dm located at height h will then experience an upward force of intensity
r g dm opposed to its weight. Utilizing the notions of friction between solids
which were introduced in Section 2.2.1, we may anticipate one of two phenomena
depending on the forces of friction prevailing around the perimeter of the wall:

• If the normalized acceleration r imparted to the sheet is not enough to over-


come the forces of friction, the sheet dm will remain compacted.
• In the opposite case, the sheet will be able to move toward the top, provided
all upper sheets are themselves free to move. If so, the material is said to
undergo decompaction.

In short, for sufficient accelerations-obviously corresponding to r greater than


I-we expect to see the formation of zones where the forces of friction exceed the
force rg dm and others where they do not. To derive the decompaction threshold
defining the boundary between the two zones, we use the maximum value of the
normalized acceleration, which is given by r = au} / g if the motion is sinusoidal
of amplitude a sin(wt).
With these assumptions, the condition that allows a sheet to move freely with
respect to the wall is simply

rgdm - gdm = dFfrict, (3-16)

whereg dm is the weightofthe sheet considered. Its mass is given by dm = pA dh,


where A is the surface area of the sheet and p its volumetric density. The elemental
frictional force dFflict can be written as dFfrict = KlhsPpv dh where P is the
perimeter of the sheet and Pv is the vertical pressure, consistent with the notation
of (3-11). The force d F flict opposes upward movements and is directed toward the
bottom.
3.2 Dynamic Properties of a Granular Pile 89

Equation (3-16) enables us to define a threshold height h t below which the mate-
rial is not allowed to separate from the wall and must therefore remain compacted,
while it will undergo decompaction above hi. The equation defining hi can be
recast in the form

The pressure Pv is itself given by (3-11), rewritten here

Pv(h l ) = pg_A_[l_ exp (-KfLs~(hl - h o))],


PKfLs A

where h o is the total height of the stack. Combining the last two equations yields
an expression giving hi

(3-17)

If we introduce once again a parameter So = P hoi A as we did in Section 3.1.4,


we may define what we call the rate of decompaction a as a = h t I h o. It is given
in terms of the normalized acceleration r by

hi In(2 - r) In(2 - r)
a=-=l+ =1+ . (3-18)
ho SoKfLs X

The rate of decompaction a turns out to depend only on the dimensionless pa-
rameter X (aside from the normalized acceleration r). This is the reason why we
chose earlier to give X the name decompaction parameter.

Levitation of Cylindrical Stacks


We are now in a position to calculate the acceleration required for levitation (syn-
onymous with decompaction) ofthe entire stack. It is called the lift-offacceleration
rio, and is given by the previous equations by letting hi = 0

(3-19)

This equation predicts that the normalized acceleration required to induce levitation
of an infinitely high stack is precisely equal to 2.

Levitation of Two-dimensional Stacks


As indicated earlier, the decompaction factor X of a two-dimensional stack is given
by X = 2hKfLsi L (see Section 3.1.4), where L is the frontal width of the structure
and h its height. With the help of the generic model developed above, it is possible
to generate a three-dimensional phase diagram of the rate of decompaction as a
function of normalized acceleration and decompaction parameter. An example
of such a diagram is presented in Figure 53. The diagram suggests a method to
experimentally test the model, which is discussed next.
90 3. Fluidization, Decompaction, and Fragmentation

1.0

i 0.8

~
0-
c:
o
0f$
<1\ 3
0-
E
o
o

FIGURE 53. Two-dimensional phase diagram of a granular stack. The parameter a is


plotted against the normalized acceleration [' and the decompaction factor X.

Experimental Observation ofDecompaction and Convection


in a Two-Dimensional Granular Structure
Simulating the dynamic properties of a granular stack by means of a two-dimen-
sional uniform structure is no doubt a bit simplistic when reality consists of a
three-dimensional pile of dissimilar particles. Nonetheless, for the purpose of
understanding the peltinent physics, it makes sense to spend the effort studying
such simple systems. There are several reasons for this:

• Two-dimensional stacks can be observed directly, as we shall see shortly. By


contrast, real three-dimensional stacks are opaque and not easily amenable
to visual characterization. To be sure, nuclear magnetic resonance has pro-
duced interesting results that, incidentally, are in substantial agreement with
the results to be presented shortly.21 But the technique is rather cumbersome,
expensive, and of limited spatial resolution.
• Two-dimensional stacks are actually the only ones that can be readily mod-
eled by computer simulation, given the limitations of current machines (see
Chapter 6).
• Experiments conducted with two-dimensional stacks have for the most part
revealed the same phenomenology as in real three-dimensional piles made,
for instance, of dry sand. This includes phenomena of convection, pile forma-
tion, segregation (see Chapter 5), and flows of various types (see Chapter 4).
That does not guarantee, of course, that the physics of two-dimensional struc-
tures is in every respect identical to its three-dimensional counterpart. There
may well be some scaling laws that could affect certain phenomena when one

21 Thetechnique makes use of medical imaging technology that has become relatively commonplace in
recent years. One drawback is that it requires particles lich in resonant radicals, most typically water.
Mustard or poppy seeds are the most popular choices of materials to work with. Unfortunately, their
mechanical properties are not very well defined.
3.2 Dynamic Properties of a Granular Pile 91

of the dimensions shrinks to small values. In particular, the organization of a


two-dimensional lattice (which exhibits translational symmetry) is likely to
favor long-range interactions.

Experimental Technique: Image Processing


Observation methods commonly used to study two-dimensional structures rely
on a variety of modern image processing techniques. Their complexity depends
on the intended goal and the imagination of the user. 22 They fall into two broad
categories.
Measurements of the Velocity ofMoving Particles
This technique makes it possible to obtain snapshots such as the one reproduced
in Section 3.2.2 [61]. It requires a stroboscopic flash that can be synchronized
with some characteristic time of the system, for instance, the excitation period of a
vibrating plate. The duration b. T of the flash is adjusted in order for the particles
to leave a visible trace whose length is proportional to their speed at the time of
the illumination. b.T must be sufficiently short to avoid random collision events
which would interfere with the interpretation of the record. Images are stored
separately and processed with statistical packages designed to compute the mean
values of the relevant quantities. It is not uncommon to be dealing with a few
thousand images, which requires automated data processing. We might note here
that this technique applies to short-duration phenomena (short compared to the
excitation period). We will see several applications in what follows.
Measurements of the Relative Motion ofParticles
Here the objective is to obtain a snapshot of a granular medium in which a number
of particles can move relative to one another. The technique is useful, for instance,
for visualizing phenomena of decompaction or convection in a stack subjected to
a vertical vibration. It works as follows:

• An image is recorded at time t), either in transmission (with backlighting) or


in reflection (front or side lighting) with a CCD camera.
• The image is digitized, meaning that it is processed into an array of zeros (dark
spots) and ones (light spots). With a suitable choice oflighting and brightness
threshold, it is possible to highlight only the centers of the particles whose
motion we wish to study.23
• A Boolean operation is then applied to a set of two images-one recorded at
tl and another one recorded previously at time t) - b.t.

220ne technique is based on Hough's criterion. It relies on numerical processing to extract the positions
of particles from very noisy signals [64].
23This is a trickier problem than appears at first glance. For instance, it is often necessary to light up
the scene from the side, in which case a correction algorithm must be used to infer the actual position
of the centers of the particles.
92 3. Fluidization, Decompaction, and Fragmentation

The truth table of an "or" operation reads:

Image I]
o I1
IImageK-
h 1 tHE
1 1

This table tells us that the resulting display would show not only the points that have
not moved between the two shots, but also the traces of those that did move. That
is how the images of convection reproduced in a later paragraph were produced.
With an "exclusive or" (or) operation, the image would show only the traces of
the particles that had moved, which may also be of interest.
The sequence of operations just described must obviously be repeated over
a sufficiently long time for something interesting to show up. In other words,
the technique is suitable to examine longer-term phenomena (compared to the
excitation period). This technique, known as "Computer Posed Photograph," or
CPP for short, was used to produce the images shown in Figure 54.
We now describe the experiment that was used to generate Figure 54. Roughly
50 x 50 aluminum beads are stacked in the form of a regular triangular lattice inside
a two-dimensional cell characterized by a shape factor So = 2h o/ L. The outer
surface of the beads is roughened by means of an appropriate surface treatment.
This is accomplished by shaking the beads for some time in air. The work hardening
resulting from the many collisions increases the coefficient offriction between balls
from an initial value of 0.2 to about 0.6. We recall that the coefficient of elastic
restitution s of aluminum balls is also about 0.6. We showed in Section 3.2.2 that
the reduced variable describing the dissipation of the collisional energy in such a
configuration is given by X z = Nz(l-s). Under the conditions of this experiment,
we are therefore definitely in a condensed regime since a column six balls high is

FIGURE 54. Computer posed photographs of a two-dimensional vibrated stacle The photo
on the right is a magnification of the region marked at the left by a circle. Convection rolls
generated by shearing at the wall are clearly evident (after [67]).
3.2 Dynamic Properties of a Granular Pile 93

enough to reach a value of X z = 3. Stated in plain language, there is no possibility


here for fluidization.
Although our configuration started initially with a flat upper surface, the exper-
iment reveals the initiation of a convection current, clearly apparent in Figure 54,
which tends to transform the stack into a pile. We should point out that, contrary
to some of our earlier claims, the phenomena illustrated in Figure 54 imply that
different parts of the stack can move relative to each other. What we are witness-
ing here are processes of decompaction and convection, which are visible in the
vicinity of the lateral walls. This is an opportune time to make a few pertinent
comments about the simultaneous existence of convection rolls and decompaction
by lateral translations within the bulk.
It is important to appreciate that the granular material we are dealing with has
a number of properties characteristic of a crystal lattice. For instance, a two-
dimensional uniform stack tends to decompact along certain directions favored by
the symmetry of the structure. Displacements occur preferentially along lines that
are parallel to or 60° off the horizontal. Taking into account the relative motions
of the lateral walls and the stack itself, which generate vertical shearing forces,
it easy to see that the resulting convection stems from slip lines oriented at 60°.
Furthermore, given the overall vertical symmetry of the structure, it should come
as no surprise that the convection rolls appear alternately to the right and to the left.
Thus, collective motions at 60° angles are necessarily accompanied by horizontal
displacements, as can be seen in Figure 54. In other words, decompaction and
convection go hand in hand, at least in this particular idealized experiment. We
should also emphasize that the phenomena discussed here occur over long time
scales and describe the stack after it has relaxed back to a state of equilibrium. As
we have already pointed out, following each vertical perturbation, the stack returns
to an ordered and energetically favorable configuration. As we will see shortly, a
totally different picture emerges when the dynamics of the stack is examined over
short time scales, in which case nonrelaxed states of the system in vibration can
be observed.
We proceed next to take a closer look at two phenomena. The first is the dynamics
of convection, which is responsible for what is referred to as pile formation in the
granular material. The second is the mechanism of decompaction, which we will
treat in terms of a continuous medium model, as was done in Section 3.2.3.

Convection and Pile Formation


The dynamic and static properties of a granular stack are particularly sensitive to
the characteristics of the container walls. That is one of the direct consequences
of the principle of redirection of stress toward the walls. The phenomenon of pile
formation in a vibrated granular material is particularly intriguing and has long
been misunderstood. It was originally reported by Faraday [12], and rediscovered
recently [65], [66]. Briefly, when a granular material is vibrated vertically, it
evolves in such a way as to develop an upper surface that is not horizontal but,
rather, inclined at an angle that is close to what is known as the angle of repose
(see Chapter 4). When the excitation is perfectly vertical and uniform, the material
94 3. Fluidization, Decompaction, and Fragmentation

takes on a shape that resembles somewhat a Chinese hat. Actually, experiments


aimed at exploring this phenomenon are not nearly as trivial to implement as it may
seem, because there exist several potential perturbations apt to obscure the primary
effect of interest. Perhaps the most troublesome among those are nonuniformities
in the excitation which can cause the particles to pile up in patterns related to the
gradients of vibration energy, much as those that give rise to Chladni's figures
discussed in Section 2.1. It appears that Faraday's original experiments may well
have been tainted by such effects.
That said, what we are about to describe applies even when the utmost care
has been exercised to subject a granular material to a perfectly uniform excita-
tion, that is, one where each point within the container experiences a displace-
ment of similar amplitude and direction. Under such conditions, the structure
adopts a conical or triangular shape (in three or two dimensions, respectively),
preserving the symmetry common to the container and the excitation. The phe-
nomenon is illustrated in Figure 55(a), which shows a CPP photograph (see Section
3.2.3) of a two-dimensional pile. It was obtained with the following experiment
[67]:

• A stack is prepared with an initially fiat upper surface, and the vibration is
started with an acceleration [' > 1.
• Convection rolls are spontaneously initiated in the upper comers of the stack,
which transport granules residing in the top portion of the structure, giving
rise to two mounds which are quite evident in Figure 55(a).

(a)

(b)
r >1
Vibrations t
FIGURE 55. (a) Photograph of a "Chinese hat" in the process of developing. (b) A simple
experiment demonstrating the crucial role played by lateral walls in pile formation.
3.2 Dynamic Properties of a Granular Pile 95

• Fairly rapid at first, the process slows down (according to a dynamics we


will study later on), as the constant addition of granules causes the two peaks
to gradually move together toward the center. The effect is generally quite
noticeable in two dimensions, and much less so in three. The end result,
however, is the same in both cases, namely the formation of a pile shaped like
a Chinese hat.

A simple experiment can demonstrate the decisive role played by the lateral walls-
more specifically, the friction between particles and walls-in the mechanism of
pile formation. It is depicted in Figure 55(b), and goes as follows:

1. The experiment uses a special container made of two concentric glass cylin-
ders, the radii of which differ by slightly more than the diameter of oxidized
aluminum pellets that make up the granular medium. Filling the space be-
tween the two cylinders with those pellets creates a two-dimensional stack
that can move freely and lacks any lateral boundary.
2. When such a structure is subjected to a vertical vibration of amplitude such
that r > I, we observe that the upper free surface remains horizontal. There
is no evidence whatsoever of pile formation.
3. With the vibration uninterrupted, a cylindrical rod of thickness comparable
to the diameter of the granules can be inserted in the space between the two
concentric cylinders, where it acts as lateral boundaries for the stack. This
immediately causes the appearance of pile formation, similar to what was
observed in the cell shown in Figure 55(a). What is more, the moment the
stick is removed, the surface returns to horizontal.
4. If the same experiment is repeated with frictionless granules, for instance,
granules with a specular polish, no pile forms whether the rod is in or out.

This simple two-dimensional experiment demonstrates conclusively that convec-


tion processes in the vicinity of the lateral walls are directly responsible for pile
formation.
Similar experiments using three-dimensional geometries confirm the key role
played by the walls of the container. In one set of experiments, illustrated in
Figure 56, two types of granules were used-some transparent and some painted
black [20]. As long as the stack is not too thick, it is possible to visually track
the path of the black granules while the cylindrical container is given a series of
vertical impulses. Figure 56 shows that the particles situated near a roughened
wall creep down along that wall. No convection is observed if the wall is smooth,
in agreement with our previous discussion.
All these observations support the contention that walls do indeed playa deter-
mining role in the mechanism of pile formation. The Janssen model also implies
that the parameter governing the onset of decompaction-and, by extension, pile
formation via convection-is the acceleration r. This has been abundantly con-
firmed experimentally both in two and in three dimensions.
96 3. Fluidization, Decompaction, and Fragmentation

(0)

FIGURE 56. Three-dimensional convection experiments. In configuration (a), the left wall
is polished smooth. Only the right wall, which has been roughened to intensify friction,
induces convection movements. In configuration (b), convection takes place in a direction
opposite to that which is observed in a cylindrical container (after [20]).

Threshold ofPile Formation and Decompaction


We report here some recent results pertaining to three-dimensional configurations
[66]. Experiments were done to determine the minimum amplitude A required for
the onset of pile formation in stacks of glass beads of various diameters subjected
to a range of frequencies. 24 The plot shown in Figure 57 demonstrates convinc-
ingly that the acceleration [' does indeed control the onset of decompaction and
convection, in agreement with the model. It must be pointed out here that these
experiments involved relatively inelastic beads which were stacked sufficiently
high for our previous arguments regarding the control parameter (Aw 2 or A 2( 2 )
to apply unambiguously (see Section 3.2.1).

Dynamics of Pile Formation in Two Dimensions


A detailed analysis-either by visual tracking or with the help of image process-
ing-ofthe mechanism ofpile formation with an acceleration slightly above thresh-
old reveals some interesting dynamic properties. As already pointed out, near
threshold, pile formation results from intermittent convection rolls that transport
(much like an Archimedes screw) some quantity of matter, namely one or more
beads, toward the summits of two emerging mounds. These ultimately merge to-
gether in the shape of a Chinese hat. A more careful analysis of the phenomenon-
over a period of several hours or even days-gives a clear picture of the dynamics

24The definition of threshold needs to be clarified. It is possible to observe slight movements of the
particles on the surface before convection is truly initiated. The ambiguous nature of this definition,
as well as other factors (such as degree of disorder, whether the stack had initially settled or not,
etc.), probably explain why these particular experiments gave a value of the acceleration r of about
1.2 at threshold, while subsequent work with two-dimensional structures gave a value of 1, within a
measurement uncertainty of ±O.05.
3.2 Dynamic Properties of a Granular Pile 97

..
Bead diameter

0 ... 0.2mm
.........
Q) 0 0 0.4 mm
"'0 0
~
::l 0 0 1.0 mm
c.. C\a
E
« Co
'--"
O'l
0
-1 '"
oOe
\
0
D
°8
,0
0

-2
1.0 1.2 1.4 1.6 1.8 2.0 2.2
log(m/2n)
FIGURE 57. Log-log plot of the vibration's amplitude threshold for three-dimensional pile
formation against frequency, for several bead sizes. The diagram shows that the acceleration
is indeed the determining factor, since the straight line has a slope of -2 (after [66]).

of the process. Short of an elaborate theory, we can at least devise an ad hoc


explanation of what goes on. We start with some results presented in Figure 58.
Experiments show a gradual slowing down of the process of pile formation,
approximately described by a logarithmic dependence on time over at least four
decades. 25 This effect appears due to increasingly less frequent occurrences of
convection rolls initiated at the walls. We note, incidentally, that this effect is
consistent with the model presented in Section 3.2.3, as well as with some ob-
servations to be presented later concerning the gradual decompaction of a stack
subjected to vibrations. Since convection rolls can exist only in a decompacted
medium, it stands to reason that they should occur less frequently as the side edges
of the Chinese hat reach down further toward the compacted zone of the pile. What
is not obvious is why this should result in a log(t) dependence.
It is instructive to derive a differential equation governing the process in order
to identify the important parameters. The empirical law x = C log(t) + B can be
differentiated to give the speed of pile formation

- x) = aexp (-cx) .
dxdt = Cexp (B-C-
25We will later on encounter another phenomenon with a logCt) dependence, for which we will propose
a more detailed explanation (see Section 4.2.1).
98 3. Fluidization, Decompaction, and Fragmentation

en
"0 25
ro
Q)
.0
.....0
r = 1.39
20
0
5 15
en
~
ro
Q)
0- 10
.....0
c:
0 5
:;:::;
'iii
0
a.. 0
1 2 3 4 5
10 10 10 10 10
Time (seconds)

FIGURE 58. Separation between each of the two mounds and the nearest wall plotted
against time for two accelerations. The experiment covers about 30 hours. Distances are
expressed in numbers of particles (after [67]).

The variable x is a measure of the distance from the summit of each of the two
mounds to the wall. Given the geometry of the system, it is also proportional to the
depth of the Chinese hat. In the above equation, the coefficient a, which describes
the efficiency of the pile formation process, turns out to depend on the acceleration
r through the quantity r - 1, and the coupling via friction between beads and
walls. 26 The coefficient C which figures in the exponential defines a characteristic
length in the experiment; it can be assimilated with the typical dimension of a
convection roll, involving the coupling between beads. In the absence of any solid
theoretical basis, it would be hazardous to venture any further interpretation of
what are after all purely phenomenological parameters. We simply conclude by
listing the experimental values of the parameters a and C, in units of number of
beads, for two values of the acceleration r.

r 1.15 1.39
a 3.9 16.8
C 1.0 2.7

Experimental Verifications ofthe Decompaction Model


With the help of the image processing techniques described previously, it is pos-
sible to test the validity of Janssen's generic model described in Section 3.2.3.
Experiments yield results that are in complete agreement with the model [63]. The
results in question are reproduced in Figure 59, which bears further comments.

26No pile formation occurs with frictionless balls or walls, or a fortiori if there are no walls at all,
as was the case in our earlier two-dimensional experiment. The same will prove true of gradnal
decompaction, to be discussed later on.
3.2 Dynamic Properties of a Granular Pile 99

1.0 '.

~::~I
t$ 0.8 " (e) O•
hh' :•..
.
0 0 .6 Decompacted t
a= hl/h o
~ 004
I.L 0.2 Compacted
I 10 So =0.67
0.0 -I-----_-----"~--o- __1
1.0 1.0 1.2 104 1.6
0.0 0.2 004 0.6 0.6 1.0 1.2
So

0.6 1.0
(b) (d)
EOA 0.8 '"

~
Decompacted
t$
0.6 '.
S
L..
0 So= 1.26
.c 0.2 t5 004
<1 ~.. 'Kf=011 CIl
I.L 0.2
Compacted
10
0.0
a 4 0.0 I
6
1.0 1.2 104 1.6
h (cm) Ie

FIGURE 59. Experimental results confirming the gradual decompaction model (after [67]).
See text.

The experimental two-dimensional cells used in these experiments are entirely


similar to those described at the beginning of this section.

• Figures 59(c) and 59(d) show the results of experiments intended to test
the relevance of the phase diagrams (a, r, X) discussed previously. The
experiments consisted in filling a two-dimensional cell with pellets of oxidized
aluminum, and placing the system on a vibrating plate whose acceleration
r = au} / g was adjustable. As predicted by the model, the experimental data
points line up with the theoretical curves corresponding to a value of the
product Kf (where f designates the friction coefficient) equal to 0.29. At
least that is the case for the two particular stack heights investigated. We
may recall that the triangular stack model described in Section 3.1.3 gave a
value K = 0.58. Since the friction coefficient fpw between pellet and wall
can vary from one cell to another between 0.2 and 0.5, we should indeed
expect the product Kf to range from 0.1 to 0.3 depending on the quality of
the surfaces of both the walls and the aluminum pellets. This turns out to be
quite consistent with experimental results. We should also note that, as the
pellets and the walls wear out during an experiment, the coefficient of friction
may actually change with time. This warrants some care in the interpretation
of the results .
• Figure 59(a) shows the lift-off acceleration rIo (equal to 2 - e- X ) as a function
of the height of a stack in a particular cell. This experiment is relatively easy
to do. It consists in measuring the acceleration required for the lower row
of the stack to lift off, which can be monitored with a CCD video camera
with a fairly high image magnification. In agreement with the model, the
entire lower row turns out to lift off in unison, providing an a posteriori
justification for simulating the stack as a superposition of sheets. In fairness,
100 3. Fluidization, Decompaction, and Fragmentation

(a) (b) (c)

Pile Formation No Boundaries Dual Lattice


K>O !l>0 K>O !l::0 1<::0 !l >0

FIGURE 60. Synopsis of results consistent with the gradual decompaction model. Three
configurations correspond to different combinations of the parameters K and /-L (after [97]).

we will see later that this behavior holds only in cells of relatively small
lateral dimensions. Section 3.2.5 will describe what happens in larger cells. As
predicted by the model, the acceleration needed for collective lift-off of a stack
is indeed a monotonically increasing function of its height. The data points fall
on a theoretical curve with a single adjustable parameter Kf = 0.11. These
results were obtained with a different cell than the one used in the previous
experiment.
• Figure 59(b) refers to a completely different experiment, although, this time,
it is done in the same cell as above. Stimulated by the vibrations, the pellets
can move relative to the cell. Provided that the particles leave a record of their
movements on the front and back windows of the cell, we can then calculate
the amplitudes of the displacements /')"h(h) as a function of height h. Here
again, there is excellent agreement between theory and experiment.

Figure 60 brings together a series of contrast-enhanced images of granular stacks


viewed through the transparent windows of various two-dimensional cells. It is
a fitting conclusion to this paragraph devoted to experimental verifications of the
model in that it summarizes all the results obtained thus far.
Inasmuch as the model identifies the quantity X = 50 K/h as the key parameter,
we have, for one thing, been able to verify the effect of the aspect ratio 50 when K
and /h are both nonzero (Figure 60(a)). The two-dimensional cylindrical container
enabled us to explore the case K > 0 and /h = 0 (Figure 60(b)).
Finally, the case K = 0 and /-L > 0, depicted in Figure 60(c), requires a trick
based on the discussion in Section 3.1.3. In that section, we analyzed how the
surface occupied by a horizontal row of parallelograms changes as a function of
a vertically applied stress. We pointed out that the dilatancy properties of such
an object vanish when the disks lined up vertically come into contact. If we now
consider a two-dimensional stack of such parallelograms in the same configuration,
it is clear that the vertical lattice, named dual lattice, must have a coefficient K
such that K = O. This configuration is illustrated in Figure 61. As it turns out,
such stacks exhibit sufficient stability to withstand moderate accelerations r in
3.2 Dynamic Properties of a Granular Pile 101

Triangular Lattice Dual Lattice


K>O K=O

FIGURE 61. Stacking in a triangular lattice and its dual counterpart, rotated by 90°. The
configuration on the left is the only one that redirects stresses toward the walls.

the range of 1 to 1.5, which are typical of the experiments conducted previously
with conventional triangular lattices. Unlike the conventional case, though, and
in full agreement with the model, there is here neither gradual decompaction, nor
convection, nor pile formation. In other words, the meshed character of such
structures is indeed one of their fundamental properties explaining not only their
dilatancy properties but their resistance to decompaction under vibrations.
Short- Term Decompaction: Fragmentation
Although the model described in the previous paragraphs is based on the mechanics
of continuous media, it accounts fairly satisfactorily for the gradual decompaction
properties of granular stacks subjected to vibrations. Before proceeding any fur-
ther, it is essential to appreciate the deeper implications of this theory. The model
rests fundamentally on Janssen's hypotheses, supplemented by the dynamical law
expressed in (3-16), which was arrived at empirically. Our purpose is not to reex-
amine here the derivation of Janssen's equation (see Section 3.2.3), but we do need
to take time out to reflect on the nature of the approximations underlying (3-16),
which is rewritten here for convenience:
fg dm - g dm ::: dFtric!.
We will focus on two key points: (1) the assumption that the system can be de-
composed in uniform horizontal sheets; and (2) the simplified description of dry
friction .
• To begin with, we recall that the model assumes that the material is made
of horizontal sheets of mass dm, rubbing against the walls where a fric-
tional force d Ffrie! is developed. This force can be calculated with the help
of Janssen's hypothesis. Furthermore, (3-16) implies that the acceleration
fg imparted by the vibrating plate is homogeneous throughout any hori-
zontal slice of material. Assuming that acceleration, mass-and, therefore,
friction-are all uniform in a given sheet is manifestly in flagrant contradic-
tion with our earlier description of the network of contact points within a
granular material (see Section 1.2.2). The reader will recall that we invoked
102 3. Fluidization, Decompaction, and Fragmentation

arches anchored onto the lateral walls to describe the equilibrium forces in
such materials. The notion of arch is an essential and unavoidable component
of the static and dynamic properties of granulars. From this vantage point
alone, we can expect substantial deviations between theory and reality.
• The extremely simplified way the friction forces with the walls are treated
may have even more serious consequences. Without going into details that
will be discussed in greater depth in Section 6.4, we merely point out here that
(3-16) expresses nothing more than a simple rupture of the forces of contact at
the boundaries when a sheet experiences a sufficient force. This completely
neglects the indeterminacy of forces which was discussed in Section 3.1.1,
and assumes that these forces are fully mobilized and directed straight toward
the top, in accordance with Janssen's hypotheses at equilibrium. It is highly
likely that the random nature of the contact chains, for one thing, and the
more or less elastic coupling between particles involved in arch formation,
for another, will in fact lead to a rather large range of contact forces at the
walls. Invoking elementary uniform horizontal sheets to model a medium that
is inherently inhomogeneous and partly undetermined may seem fraught with
danger. Furthermore, (3-16) also ignores the coefficient of dynamic friction
and, perhaps more importantly, the velocity dependence of the friction force.

All these considerations may induce the skeptics to think that the agreement be-
tween theory and experiment is purely fortuitous. That conclusion would be
wrong, though, at least as long as the model is used with proper caution and within
its own limitations. It is appropriate to point out, for instance, that the simula-
tion of gradual decompaction merely establishes a phase diagram that predicts
the average height at which a stack undergoes decompaction, without specify-
ing anything further about what might be going on in the decompacted phase.
Except for the calculation of particle trajectories, which requires an additional
assumption to be discussed shortly, the model does not go beyond working out
the height of the compacted phase. The procedure usually takes into account the
irregular character of Coulomb's friction law, as we will discuss in more detail in
Section 6.4.
The experiments we have described thus far essentially deal with long-duration
observations. In actuality, the decompaction processes observed experimentally
proceed by way of successive relaxation states in which a system remains long
enough (typically several seconds) to be recorded by conventional techniques, such
as CPP photography. This approach misses the details of any fast event (lasting
no more than a fraction of the excitation period) involving short-duration and re-
versible distortions. While that much is clear from an experimental point of view,
what consequences does it entail in terms of the simplified model we have devel-
oped? In what fundamental way is it-in spite or because of its imperfections-
more suitable to account for long-duration phenomena? There is no clear answer
to this question. All we can do at this point is to simply acknowledge that the model
does seem to describe fairly accurately systems in their reorganized states, after
they have relaxed back to equilibrium. It does not address the largely indeterminate
3.2 Dynamic Properties of a Granular Pile 103

Fracture
Lines

FIGURE 62. Short-term observations, using a strobe light synchronized with the vibration
period, reveal that decompaction of a vibrated two-dimensional pile actually results from
a series of fragmentations propagating from top to bottom during ballistic flights. Such
fragmentations occur during each cycle of the excitation (after [69]).

solutions, subject to random fluctuations, that would result from a more complete
analysis of the problem.
Given the fundamentally dual nature-short-duration and long-duration-of
the relevant phenomena, it is naturally tempting to observe the time evolution of a
granular stack by imposing a slight shift in the frequency of the stroboscopic light
flashes relative to that of the excitation. Figure 62 shows the results of just such
an experiment. 27
The image contains a lot of information about the short-term behavior of a
granular structure, which deserves a number of comments. 28 First of all-and
that cannot be conveyed by a frozen snapshot-there are fractures, such as the
one indicated in the figure, that have dynamic properties of their own. A video
would show that these fractures are initiated near the top and propagate toward the
bottom. This mechanism of successive fragmentation occurs during each cycle
of the excitation and persists during the ballistic phase of the stack, which can be
quite short compared to the excitation period (typically in the ratio of 10 ms to
0.2 s). These fractures have a characteristic V-shaped form (in the present case
pointing toward the bottom) and are more or less dislocated as they progress within

27 This discussion does not respect the actual chronological sequence of events that led to this exper-
iment. In reality, it was while trying to measure directly with a video camera the movements of
particles in a granular stack that researchers came to realize that gradual decompaction resulted from
a series of more or less irreversible fragmentations.
28Note that CPP images and traces left on frontal windows can only provide information averaged over
large numbers of periods.
104 3. Fluidization, Decompaction, and Fragmentation

the structure. We shall see in the next section how we can extract more detailed
information about this phenomenon, which appears to be a very general mode of
decompaction of a guided stack in ballistic flight.

3.2.4. Fragmentation of a Stack in Guided Fall


Having observed that gradual decompaction actually results from a succession of
fragmentations that separates the initial pile into smaller-size blocks, it becomes
logical to look for an experimental configuration that would make it easier to study
and quantify this phenomenon. As it turns out, trying to isolate the process of frag-
mentation can be singularly complicated by the configuration of the stack under
vibration. The reason is that, as we have indicated on several occasions, what
drives decompaction is the friction between particles and walls. In the preceding
experiment, this shearing interaction stemmed from the differential motion of the
granular material-in more or less parabolic ballistic flight-relative to the con-
tainer walls, whose movement is sinusoidal. On the other hand, direct observations
show that fractures appear rather randomly from one period to the next. These
fluctuations, which are of considerable interest in their own right, unfortunately
seriously complicate the analysis of the elementary mechanism to be studied. The
experiment described next was devised in an attempt to simplify as much as pos-
sible the conditions leading to the observation of this phenomenon [69].29

Two-Dimensional Experiment
Particles are stacked in a regular triangular arrangement in a vertical cell of the
type described previously. The bottom of the cell is blocked by a spring-loaded
metal plate that can be dropped rapidly (with an initial acceleration of 3g). A
CCD camera records the fall of the stack, which lasts approximately one-tenth of
a second. Figure 63 shows a time-lapse sequence obtained during the experiment.
As expected in light of the results presented in the preceding paragraph, the stack
breaks up into separate V-shaped blocks whose apex now points upward, always
in a direction opposite to gravity. 3D Also as noted before, successive fractures
propagate toward the top of the pile during the fall. All these observations are fully
consistent with those described in connection with a vibrated stack, provided that
the sign of the acceleration r imparted to the structure in ballistic flight be reversed.
While it was directed toward the top in the case of a vibrated cell (upward ballistic
flight prior to recompaction), it now points down. Given our previous discussion
of the formation arches (see Section 3.1.1), the V-shaped structures observed in the
present experiment can be understood as contact chains with enhanced strength
supporting the granular mass above them.

29Drake reported experiments involving inclined falls in which he was studying structural modifica-
tions by direct observation, although he did not draw the conclusions derived here concerning the
fragmentation process [68].
30 Similar observations apply to three-dimensional structures. If a small tube about I cm in diameter is
half filled with compacted sand and the bottom is suddenly opened, the content escapes by fragmenting
from the bottom up into a series of blocks separated by arches analogous to those shown in Figure 5.
3.2 Dynamic Properties of a Granular Pile 105

FIGURE 63. Experimental observation of fragmentation of a two-dimensional stack in


guided fall. The photographs are taken every loth of a second. A fragmentation wave is
seen to propagate toward the top as the pile drops down the tube. One particle has escaped
at the bottom and falls in accordance with gravity's acceleration, while the top of the stack
falls with a lower acceleration (see text). The stack decompacts by fragmenting (after [69]).

Theoretical Modeling
An interesting question is whether our previous theoretical model, which is a dy-
namic extension ofJanssen's model, can account for the process offragmentation. 31
We will later show how a properly adapted numerical simulation can indeed re-
produce the appearance and propagation of fractures within granular piles [69].

Fall Without Fragmentation


The first question that comes to mind has to do with the fall of a simple block
assumed, for the time being, to be free of fractures. This problem was actually
solved in Section 3.2.3, and we use the same notation here, including the convention
that heights be counted positively from the bottom up. The height-dependence of
the vertical stress was given by

PvCh) = pg_A_[l_ exp


PKfLs
(-KfLs~(h
A
- ho»)].
In two dimensions, this equation reduces to

L [ 1 - exp ( --L-(h
PvCh) = pg 2KfLs 2KfLs - h o )] ,
- h o))] = pg~ [ 1 - exp (h-~-

31Savage studied the inclined or vertical fall of a noncohesive granular system [70]. Based as it
is on equations describing the mechanics of continuous media, Savage's model is essentially a
hydrodynamic description, while ours, which relies on the concept of fractures, is more akin to
models favored by geophysicists.
106 3. Fluidization, Decompaction, and Fragmentation

where, for convenience, the characteristic coefficients are lumped into a single
parameter C; = Lj(2Kfls).
A straightforward solution can be obtained, considering the stack as a whole
whose apparent weight has been given by equation 3.12. In this case and supposing
that the pressure and friction forces remain constant during the downfall process
of a pile whose height is h o, we get the reduced acceleration of the stack as

where X = ®. !;
An alternative approach would first consist in supposing that the pressure equi-
librium has been established before opening the spring-loaded metal plate. Second,
and because of the particular discontinuous nature of the material, we may con-
sider a thin slab of granulate and suppose that the previously mobilized friction
forces are still acting at the wall thereby reducing the downwards acceleration of
this particular piece of material. We know that the friction force developed at the
walls at this particular thin slab of granulate can be written as

- - = -Pv = g [ 1 -
dFfrict -- -
exp (h hO)]
dm pC; C;

Thus, a thin slice of mass dm of granular material located at height h is subjected


to an effective acceleration y (h) smaller than the acceleration of gravity. It can be
written as
dFfrict
y(h) = gf(h) = g - --,
dm
which yields the normalized acceleration experienced by this particular slice at
height h

h - ho )
f(h) = exp ( -C;- where h E [0, hoJ.

Strictly speaking, and in the context of the hypotheses discussed earlier, this equa-
tion is valid only at the precise instant when the downward fall is initiated. In the
interest of simplification, and for a semiquantitative analysis, we shall assume that
it holds for the entire duration of the fall. We will see later on that experiments
justify this approximation.
We begin by noting that r(h) is an increasing function of the height h. This
implies that a stack subject to its own weight and to the applicable forces of friction
will tend to remain compacted as it falls. In other words, the stack experiences no
spontaneous tendency to fracture, unless external causes come into play.
To identify external causes likely to initiate fractures that can propagate and
lead to fragmentation, we need to modify the surface properties of the lateral
walls. When we do so, we observe the following:
3.2 Dynamic Properties of a Granular Pile 107

I
I
Fractures
400 One particle

E
.s..9! 300
ho
ho=20 mm

"6. ho=44 mm
Q)
:S
'0 200 ho==' 93mm
Cl.
B
Q) h o= 130mm
:S
'0 100
1: ho==' 195mm
0>
"iii
J:
0

0 50 100 150 200 250


Time (ms)

FIGURE 64. Altitude at the top of the pile as a function of fall time. The filled symbols are
experimental data. The solid curves correspond to parabolas calculated with KfLs = 0.12.
The arrows indicate the appearance of at least one fracture in the pile (after [69]).

• Surfaces prepared with a specular polish, that is to say, with a roughness of the
order of a fraction of a micron, have a coefficient of friction fLs that is virtually
the same as for regular surfaces after they have gone through normal wear-
out. With polished lateral surfaces, stacks can fall without fragmentation,
although with a reduced acceleration compared to that due to gravity alone, in
accordance with (3-20). These conclusions emerge from experiments whose
results are shown in Figure 64. They were obtained with identical oxidized
aluminum pellets placed in the same cell, but stacked to different heights.
The data indicate that within experimental errors, and as long as no fracture
appears, the stack falls with a normalized acceleration that is governed by the
initial height and is adequately described by our equation for r(h). The solid
curves in Figure 64 were calculated using a single adjustable parameter KfLs
of 0.12, which is somewhat smaller than the value determined previously. The
discrepancy is probably due to the fact that what matters here is the dynamic,
rather than static, coefficient of friction.
• Lateral surfaces with a roughness of a few microns or more trigger the ap-
parently random appearance of a series of ascending fractures, as noted pre-
viously.

Where Do Fractures Initially Appear?


It is extremely difficult to accurately model interactions between a stack and the
surface of the lateral walls. Accordingly, we will restrict ourselves here to the
issue of the stability of a fracture initiated at one of the lateral walls by a random
fluctuation of the topography, or of the micromechanical parameters (which can
disrupt the regularity of a lattice), or even by a loss of contact with the wall or
any other instability apt to promote fracturing. Let us then assume that a fracture
108 3. Fluidization, Decompaction, and Fragmentation

appears at height h f measured, as usual, from the base of the pile. The fracture
causes an interruption in the contact chain and a reorganization (assumed to be
instantaneous) of the equilibrium forces in what is now two daughter piles-pile
A on top and pile B at the bottom. Each daughter pile is from now on subjected
to its own acceleration r A and r B .
Since r is a monotonic function of height, it is immediately apparent that a
fracture will remain stable or can even get amplified as long as r A ~ r B, which
implies the condition h f .:::; h al2. Stated in words, the fracture must be initiated
in the lower half of the stack if it is not to close up on itself during the fall.
We next examine how an ab initio simulation can reproduce many of the results
we have just discussed.

Numerical Simulation of a Stack in Guided Fall


We have learned in Section 2.2.2 how to write down useful equations for simulat-
ing numerically the dynamics of stacks. Still more techniques will be presented
in Chapter 6. Until then, we further our goal to understand the physics of frag-
mentation of granular stacks in guided falls by examining a particular numerical
technique which provides additional insight into a number of experimental obser-
vations discussed above.
The method we are about to use is rooted in the concept of rigid spheres (in a
sense to be specified in Chapter 6). It is an "event-driven" approach quite consistent
with the spirit of that which we introduced to treat the problem of a one-dimensional
column in Section 3.2.1. As we stressed then, the technique is fundamentally dy-
namic in nature and is not really suitable to account for static situations such as the
initial state of a stack. Its implementation presupposes the use of an artificial "ther-
mal agitation" enabling the particles to sense, as it were, their immediate surround-
ings, even in a static condition. 32 In this picture, equilibria seem themselves de-
scribed in terms of entities like impulses, linear and angular momenta, all of which
are related to translations and rotations. A real equilibrium or quasi-equilibrium
situation contains inherent uncertainties that may obscure the connection between
experimental reality and numerical simulations. As far as experiments are con-
cerned, random irregularities built in the surface of the lateral walls often set off
fractures. By the same token, a thermal agitation artificially inserted in a numerical
model constitutes from the outset a random phenomenon capable of initiating frac-
tures that are independent ofthe quality ofthe walls. That being understood, numer-
ical simulations turn out to reproduce satisfactorily many aspects of experimental
results, as illustrated in Figure 65. Without going into the details of the simulation,
which are discussed in detail in reference [71], we highlight a few salient results:

32The introduction of an artificial thennal agitation to simulate, through the bias of a dynamic numerical
technique, the static properties of a stack hides a profound reality. As we shall see in Section 6.4, the
dynamic properties of a stack are less subject to indeterminacies (in the sense of Section 3.1.1) than
are its equilibrium characteristics. Forces are mobilized and well defined at all times during multiple
collisions, whereas unknowns persist at rest. Furthennore, thennal agitation implies a fictitious
source of fluctuations which has no basis in reality. This touches on a fundamental problem in
matters of simulation and understanding of the behavior of granular materials.
3.2 Dynamic Properties of a Granular Pile 109

t =0 s 0.02 s 0.04 s 0.06 s 0.08 s ,


I

FIGURE 65. Simulation of the fracturing of a stack in guided fall under conditions similar
to the experiments. Note that thermal agitation, required by the simulation, causes a kind of
artificial boiling effect at the top of the pile. The time t is recorded in seconds. The aspect
ratio and micromechanical coefficients match the experimental values (after [69]).

• Numerical simulations produce satisfactory descriptions only if the angu-


lar momenta of particles are included. Otherwise, the results look more
like a gradual dilution of the stack, from which inverted V-shaped structures
are conspicuously missing, even though they are invariably present in ac-
tual experiments. The implication is that rotations playa key role in this
phenomenon. This point will be examined further as we go along.
• Fractures appear and grow near the bottom of the stack, in agreement not
only with observations but also with the simplified model discussed above.
As predicted, those fractures that occasionally appear in the upper portion of
the stack are quickly squelched by collapsing back on themselves during the
fall.
• Fractures typically get started on one side of the structure, violating the sym-
metry of the system. They typically emanate from a boundary, much like
what is observed in cantilevered fractures of solids. 33 This trend can also be
observed in numerous snapshots of real experiments.

Having recognized that simulations must take into account the rotation of particles
if they are to produce results in agreement with reality, we still have to understand
why that is. The best strategy to that end is to rely on experiments. That is precisely
the course of action we will follow. We will begin by considering the distribution

33 A cantilevered fracture occurs, for instance, when we attempt to cut a solid object with a knife. A
fracture often opens up at the point of contact of the knife and propagates according to a dynamics
familiar in geophysics.
110 3. Fluidization, Decompaction, and Fragmentation

0.25
t=0.02 s -+-
t=0.04 s .......

t
~ t=O.06 S .". •..
~ 0.20 Ijl
t=0.08 S ..·H....·
c
:::J
~
~ 0.15
:e
~
[1! 0.10
:::J
<Jl
<Jl
[1!
Q.. 0.05

0.00
-0.05 0.00 0.05 0.10
z(m)

FIGURE 66. Numerical simulation of the pressure against the walls as a function of height.
The integration time is 10 ms, and each point is obtained by averaging over a height
equivalent to six rows (after [69]).

of stresses in a stack in guided fall, and go on from there to examine in some detail
the modes of self-organization of the rotation of a collection of particles in the
immediate vicinity of a fracture.
Distribution of Pressure in a Stack: Arch Effects
Numerical simulations allow us to keep track of the velocities of the particles and
the rate of collisions between themselves as well as with the walls at any point
in space and time. As we have emphasized earlier, we are dealing here with a
dynamic model. As such, we are not in a position to calculate a static pressure.
Rather, we consider that the pressure exerted by the particles results from repeated
collisions and transfer of momenta to the lateral walls such as described by (2-10),
in which the normal component is conserved. This approach is entirely consistent
with the traditional kinetic theory of gases. The time integration of the momentum
I~"t P dt is carried out over a reasonable duration of the order of one-hundredth of
a second. The results of such calculations, obtained for a series of different times,
are shown in Figure 66.
The appearance offractures within the stack (at 0.04 and 0.06 s) corresponds to
a significant increase-by one order of magnitude-of the local pressure on the
walls. This is very much in keeping with our intuitive notion of an arch whose
primary function, as we have seen in Section 3.1.1, is precisely to transfer the
pressure exerted by the upper portions of the stack laterally to the walls. Thus,
through the bias of a numerical simulation, we arrive at some objective and quan-
tifiable information about the mechanism of fragmentation. It appears to result
from the successive formation and disappearance of triangularly shaped arches
pointed upward, as we may have anticipated.
It is likewise possible to study in detail what causes fractures to form in the stack.
The model suggests that a fracture opens up with a single particle whose translation
and rotation velocities (see Section 2.2.2) suddenly have to adjust to the relative
3.2 Dynamic Properties of a Granular Pile III

(b)

I""-~
I
I

(a) /\ Line of fragile contacts

(c)

Impossible Configuration

FIGURE 67. Diagram (a) shows a simulation of rotations in a stack in guided fall. The
rotations organize themselves around fractures indicated by arrows. Diagram (b) shows an
organization compatible with V-shaped contact chains of the type observed experimentally.
The arrangement depicted in diagram (c) is incompatible with the arch model and is in fact
never observed in simulations.

speed of the wall. Stated differently, unlike its cousins in the interior of the stack, a
particle interacting with a wall goes into a regime of rolling without gliding, in the
sense defined in Section 2.2.2, which triggers the onset of a fracture. Furthermore,
the model also favors the propagation of a fracture toward the interior along easy
dislocation lines of the lattice. 34 This propagation results in a dramatic increase
in the number of collisions between particles along and around the contact chain.
It would then appear that the mechanism responsible for initiating a fracture is
a kind of organization of the rotations of particles in relation to the wall. Given
the concomitant increase in the number of collisions in the immediate vicinity
of the event, it is logical to inquire whether such a local organization can in fact
propagate along a contact chain. That is our next topic.

Self-Organization of Rotations
Figure 67 depicts a map of the "spins" of particles in a stack right after the mo-
ment two fractures, indicated by arrows, have opened up at the wall. 35 White circles
indicate a counterclockwise rotation with an axis perpendicular to the plane of the
figure, while black circles correspond to a rotation in the opposite direction.

34We use here the standard terminology in crystallography. Given the triangular symmetry of a compact
two-dimensional stack, the dislocation lines run horizontally or at a 60° angle with the horizontal.
35 In the present context, the word "spin" designates the rotation of a particle with an angular momentum
perpendicular to the plane of the two-dimensional stack.
112 3. Fluidization, Decompaction, and Fragmentation

Figure 67(a), obtained by numerical simulation, shows clearly that the spins
tend to organize themselves in an alternating pattern in the immediate vicinity of
and just above the fractures. Figure 67(b) and (c) may help us understand that the
angle of these patterns is compatible with the notion of arch developed earlier in our
discussion. A contact chain inclined at a 60° angle relative to the horizontal implies
that the rotations of its strongly bound particles alternate. It is intuitively obvious
that the particles forming such a contact chain, being in intimate contact with each
other, will experience frustrated rotation relative to the chains immediately above
and below (see Section 2.2.1). It follows that the weakest points of the structure
run between superposed contact chains-parallel to the direction of the arch-and
that fractures can develop only along those lines. Here again, we see that the
self-organization of the rotations of particles conforms to the notion of arches as
contact chains which we developed earlier. 36

3.2.5. Surface Instabilities in an Extended Granular Medium


Many of the phenomena we have described thus far, such as convection or frag-
mentation, can be traced to interactions between granular media and the walls of
the containers. Before moving on to the topic of free inclined flows in Chapter 4,
we end this chapter with a variation on the previous case, where the stack is now
much wider than it is high. It is useful to recall some considerations we invoked
in connection with Janssen's static model (Section 3.1.4) and its dynamic exten-
sion (Section 3.2.3). We pointed out that, within the constraints of the model, the
behavior of a stack of height h and perimeter P is governed by the decompaction
parameter X = S K f.-t through the exponential factor exp (- X). Recall that the as-
pectratio S = Phi A is a measure of the lateral area normalized to the cross-section
area A. In a two-dimensional configuration, the aspect ratio becomes S = 2hl L,
where L is the lateral length. The attenuation of the stresses in the stack may be
written as

exp(-x) = exp (-~) = exp (_ 2h)


LIKf.-t A '

where the quantity A == L I Kf.-t is the characteristic distance of this attenuation. It


may be interpreted as the mean height of an arch redirecting the stresses toward
the walls since its effect is to limit the vertical pressure due to the weight of a
column. In a typical case, where Kf.-t ~ 0.3, the height of an arch is about three
times the width of a silo. We may thus expect to see effects at variance with pure
hydrostatics in two- or three-dimensional silos whose heights are some two or three
times their widths. Conversely, when the height of a stack becomes much smaller
than its lateral extension, which means h «L (or h «AI P in three dimensions),

36In this picture, the formation of arches and fractures is a consequence of a process of self-organization
of rotations that gives rise to lines of easy fracture (particles in frustrated rotation) on the one hand,
and lines of energetic contacts (particles with alternating rotations). Extrapolating this picture, we
might even envision that an initially amorphous granular material may "crystallize" in sheets that are
easy to peel apart but are quite strong in their own plane. Given our present state of knowledge, this
possibility remains purely speculative.
3.2 Dynamic Properties of a Granular Pile 113

(a) (b)

A
~ Vibrations

Vibrations ~
FIGURE 68. Schematic diagrams of experiments to observe instabilities in thin granular
layers under vibration. Patterns are monitored from the top in three-dimensional configu-
rations, and sideways in two dimensions.

the phenomena we discussed in the preceding paragraphs should take on a very


different look. We proceed to explore this issue further.
Figure 68 describes two experiments conducted recently. They depict surface
patterns that develop spontaneously when extended granular stacks are subjected
to vibrations [72], [73]. Figure 68(a) descrit.;s a three-dimensional arrangement
in which a container of inside diameter d = 127 mm holds seven layers of bronze
balls whose diameters range from 0.15 to 0.18 mm. As we discussed in Section 2.1,
these objects are small enough to warrant placing the container in a partial vacuum
of 0.1 torr to avoid coupling effects with the ambient fluid. We also mentioned the
risk that any agitation might create electrical surface charges on the metallic balls.
To facilitate draining such charges to ground, the base of the container is made of
aluminum. The aspect ratio of this particular system is

Ph 4h
S == - = - Rj 0.01.
A nd
Remarkably, in this three-dimensional situation, large changes in the coefficient
of restitution (from 0.5 to 0.95), of the density (from 2.3 to 11.4), of the aspect
ratio, and of the number of stacked layers, seem to have virtually no influence on
the phenomena we are about to describe. This is not true of a two-dimensional
configuration, however. In that case, we will see that the number of layers has a
pronounced effect.
Figure 68(b) illustrates a typical two-dimensional experiment. Once again, the
cell is made of two glass plates separated by a distance slightly greater than the
diameter of the particles, which in this case are 1.5-mm aluminum beads. The lat-
eral dimension of the cell is 300 mm, and it contains a number Nh of layers
varying from 4 to 27, which corresponds to aspect ratios of a few percent. Given
the much larger size of the beads, it is not essential this time around to work in
vacuum, and electrostatic interactions are all but negligible. We may even use
114 3. Fluidization, Decompaction, and Fragmentation

(a) (b) (c)


FIGURE 69. Three typical organization patterns on the surface of a 1.2-mm-thick three-
dimensional granular pile vibrated at a frequency of 67 Hertz. Diagram (a) shows striations
at f /2 (f' = 4.0). Diagram (b) is an example of competition between squares and striations
at f /4 (f' = 6.0). Diagram (c) shows hexagons at f/4 (f' = 7.4) (after [74]).

oxidized aluminum balls without significantly affecting the resultsY Needless


to say, precautions were taken in both sets of experiments to ensure that the re-
sulting patterns were not somehow due to interactions with the lateral walls. In
distinct contrast to earlier situations, we are dealing here with instabilities gener-
ated within the mass of the material by the sinusoidal vibration of the container.
Although the overall phenomenology is essentially the same, we will discuss two-
and three-dimensional cases separately. We begin with the behavior of extended
three-dimensional stacks, which in many respects are reminiscent of those Faraday
observed as early as 1831 on the surface of vibrated liquids [12].38

Extended Three-Dimensional Stacks


When the normalized acceleration r is varied, experiments show a succession
of patterns developing on the surface of a vibrated granular [72]. The mate-
rial organizes itself in striations, squares, hexagons, as well as combinations of
those for particular values of the acceleration r = 4n 2 /2 AI g (A is the max-
imum amplitude of the sinusoidal excitation, and / is the frequency) transmit-
ted to the cell. Figure 69 shows a few examples of such patterns viewed from
above.

37There is a deeper significance attached to this observation. Our earlier decompaction model es-
tablished that in a vibrated granular material with a large aspect ratio, the behavior of the stack
was determined by the friction between particles and walls. Even though the present experiment is
fundamentally different, we nevertheless find again that friction between particles does not seem to
influence the overall behavior of thin stacks under vibrations. We might add that no one understand
clearly why that is.
38Faraday was studying the behavior of viscous liquids by subjecting their container to periodic vertical
vibrations. He observed the development of surface waves rather similar to the ones described here.
He could not, of course, have seen bifurcation phenomena, which are characteristic of inelastic stacks
(see Section 3.2.1).
3.2 Dynamic Properties of a Granular Pile ll5

3
Disorder
Hexagons

.....
..c
~ 2
'5
Ql
E
i=
Flat

o
2 3 4 5 6 7 8 9
Acceleration r
FIGURE 70. Bifurcation diagram of inelastic particles and organization patterns of the
surface of an extended three-dimensional granular medium (after [74]).

The formation of these geometrical patterns results from the combined effects
of two phenomena, at least one of which is specific to granular matter:

• First is a phenomenon of successive bifurcations associated with the vibra-


tional perturbation of a granular material behaving as a completely inelastic
object, analogous to what we encountered in Section 3.2.1. 39
• Second is a mechanism of parametric excitation of surface waves of the type
observed in a liquid (the ones Faraday observed). This phenomenon involves
a complex filtering effect-via anharmonic coupling-of the harmonics and
subharmonics of the excitation frequency.

Additional insight is provided by Figure 70, which reproduces in a slightly different


format the bifurcation diagram shown in Figure 40, combined with the results of
experimental observations relative to the formation of self-organization patterns
such as striations, squares, hexagons, and various combinations thereof.
This bifurcation diagram plots the product f . tf' t f being the time of free flight
of the granular layer, against the normalized acceleration r. It reveals the existence
of critical points separating regions corresponding to different geometrical figures.
As the acceleration r is increased, we observe successively striations at f /2,
hexagons at f, striations and squares at f/4, hexagons also at f/4, and eventually
almost complete chaos when r reaches the value 8. For frequencies between 10

39Considering how thin the layers are in this experiment, we might legitimately be concerned that the
condition N" (1- 8) :c: 3, required to ensure that the stack will not be fully fluidized, may be violated.
The reader may convince himself that eight balls with a coefficient of restitution of 0.6 are sufficient
for the condition to be verified. As it turns out, compliance with the above condition is not crucial
for this particular experiment.
116 3. Fluidization, Decompaction, and Fragmentation

30

o
.000 .002 .004 .006
2
1/f (sec 2 )

FIGURE 71. Wavelength of organization patterns plotted against the inverse of the square of
the excitation frequency. Both straight lines have a slope of 1, indicating a linear dependence
for spheres of diameter 0.4 mm (square data points) or 0.2 mm (triangular data points). Only
the ordinate at the origin depends on the size of the spheres. In this case, r = 3.5 (after
[74]).

and 100 Hertz, these experiments confirm that it is indeed the parameter r that
governs the appearance of self-organized surface patterns, as well as their contrast,
validating the model of an inelastic ball.

Wavelength Dependence on Vibration Frequency


As we have just seen, the appearance and evolution of self-organized surface
patterns are directly connected with the bifurcation diagram characteristic of an
inelastic ball. What remains to be explained is precisely what gives rise to partic-
ular geometrical figures resembling in many ways those observed with Faraday's
instabilities. For that purpose, it is instructive to examine how the wavelength A of
these figures depends on the excitation frequency f. Figure 71 shows the results
of experiments aimed at studying the issue [72], [74].
The empirical results may be expressed in the form

where Amin depends only on the diameter d of the particles and is approximately
equal to lId. A similar equation appears in the book Fluid Mechanics by Landau
and Lifshitz, in a paragraph dealing with gravitational surface waves. 40 In this
light, the parameter geff can be construed as an effective acceleration which must
be a fraction of gravity's acceleration. That is indeed the case here, since we find
2
geff ~ 3.1 m/s .

40See page 36 of L. D. Landau and E. M. Lifshitz, Fluid Mechanics (Oxford: Pergamon Press, 1987).
3.2 Dynamic Properties of a Granular Pile 117

(a)

FIGURE 72. Organization pattern of a nine-particle-deep pile under moderate acceleration


(f = 3.4). Diagram (a) was obtained at a frequency of 7.8 Hertz, and diagram (b) at 12
Hertz. The horizontal scale is different in the two cases. In reality, the wavelengths are
such that Al = 2A2.

Extended Two-Dimensional Stacks


Experiments with two-dimensional structures show essentially the same type of
phenomenology. A shallow two-dimensional stack undergoing a vertical vibra-
tion arranges itself in a periodic pattern, typical examples of which are shown in
Figure 72. Here again, experiments have been conducted at various frequencies
[73]. The results can be summarized as follows:

• At low frequencies (f < 10 Hertz), the wavelength A of the periodic pattern


can be expressed as
A G eff
-JNh = Amin Cd) + f2'

which is substantially the same as in the three-dimensional case, except for


the fact that the wavelength has been normalized to -JNh so as to make the
expression independent of the height of the stack. No such normalization
proved necessary in the three-dimensional case. Perhaps the reason is that
measurements in two dimensions tend to be more accurate. We might also
note that Amin Cd) has roughly the same value in two and three dimensions.
• At high frequencies (f > 10 Hertz), the wavelength A reaches a plateau and
no longer depends on frequency, which is a rather unexpected result.

At the present time, the reason for the existence of two distinct regimes remains
unknown, as does the -JNh dependence at low frequencies.

Summary
The self-organization of surface patterns in extended two- or three-dimensional
granular media result from the superposition of two well-known phenomena. One
118 3. Fluidization, Decompaction, and Fragmentation

is related to the bifurcation diagram associated with an inelastic ball, and the
other is Faraday's instabilities in liquids. The latter connection is intriguing and
suggests the possibility of making a parallel between the viscosity of a liquid and
the degree of looseness of a granular sheet. Experiments indicate that a loose
granular medium indeed has a lower apparent viscosity (in the sense defined for
liquids) than a compacted one (see, for instance, Section 2.4.2). Intriguing as it
may be, this type of analogy remains largely in the realm of speculation.
4
Granular Media in a State of Flow

In Chapter 3, we established that under the influence of external stimuli, such as


a vertical vibration, a granular structure can be fluidized and decompacted, and
that it can be the seat of convection rolls and fragmentation processes. We came to
appreciate the crucial role played by the container walls in transmitting collisions
to the interior of the material and in activating friction forces. The present chapter
is devoted to the dynamic properties of the free surface of a granular material when
it is inclined at a sufficiently large angle. As we mentioned before (Section 2.4.2),
when a granular pile is tilted at an angle greater than some threshold to be specified,
a flow in sheet form is set in motion. The threshold angle constitutes a critical state
of the granular structure. We will not concern ourselves here with flows through
cones, nor will we treat more complex geometries such as hourglasses, all of
which involve arches anchored to the walls and friction between particles and
boundaries. Before diving into the topic of how a granular medium begins to flow,
we will consider more specifically the equilibrium of a stack inclined at less than
the threshold angle.

4.1. A Sand Pile in Equilibrium: The Angle of Repose


We all know that with dry sand it is impossible to make a pile with sides that are
vertical or even steeply inclined relative to the horizontal. As soon as the slope
exceeds a certain value, the pile collapses (i.e., relaxes back) until the slope returns
to an angle e which, curiously enough, always seems to be near 35°. This angle
is referred to as the "angle of repose." It might actually be more accurate to speak

J. Duran, Sands, Powders, and Grains


© Springer-Verlag New York, Inc. 2000
120 4. Granular Media in a State of Flow

FIGURE 73. The figure on the left shows the embankment angle of a two-dimensional pile.
The one on the right depicts one of many ways to create a cone-shaped sand pile. Grains
are dropped one at a time on a roughened horizontal base.

of angles in the plural, for it turns out that there is more than one, as we shall see
shortly. The phenomenon is illustrated in Figure 73.
The first individual to make quantitative observations on the angle of repose was
Charles de Coulomb, who back in the eighteenth century was a military engineer
responsible for building fortifications [11]. Being well acquainted with friction
between solids, as mentioned in Section 2.2.1, he proposed a simple explanation
that is still considered the authoritative word on the topic. His model is rooted in
the idea that two contiguous sheets of a dry granular material cannot slide relative
to each other unless their inclination e is at least equal to tan-j(ILs)' By analogy
with what we know about friction between solids, ILs is to be interpreted as a
coefficient characteristic of the friction forces involved.
This analogy suggests that Coulomb's friction law applies to granular media as
well, which is indeed supported by experimental observations, as pointed out in
Section 2.4.1. Yet, for all its simplicity, this line of thinking raises a number of
questions on closer examination. For starters, how do we define the "weight" of a
granular sheet? We have already emphasized repeatedly that the distribution of
forces exerted on the surface of such a sheet is anything but uniform and a far
cry from the simple description applicable to a massive solid placed on a support.
Based on the arguments we invoked earlier concerning the microscopic mecha-
nisms of solid friction, we can easily appreciate that the problem is even thornier in
the case of a granular sheet. The many indeterminacies and hysteresis phenomena
alluded to earlier all contribute to complicating the task of defining unambiguously
what we mean by angle of repose. Indeed, both experiments and theory show that
there are not one but several angles of repose depending, among other things, on
how a pile was prepared.
A wealth of technical details on this topic can be found in the book by Brown
and Richards [5]. We will simply summarize them here by stating that the relative
indeterminacy of the angle of repose stems from two main factors:
4.1 A Sand Pile in Equilibrium: The Angle of Repose 121

Concave Convex

FIGURE 74. A convex pile generally has a lower embankment angle than a concave file,
such as exists near the opening of an hourglass .

• The first factor is geometrical, as it involves the shape or, more precisely, the
curvature of a pile. The point is illustrated in Figure 74, which may help us
realize intuitively that particles in the vicinity of the free surface tend to be
more densely surrounded in a concave pile (in the form of a crater) than in a
convex one (shaped like a mound).
Any difference is expected to vanish when the radius of curvature of the
surface becomes much larger than the mean diameter of the granules. These
predictions are essentially confirmed by experimental results such as those
reproduced in the following table:

Material Geometry "Mound" "Crater" Dynamic


angle angle angle
Tapioca spherical 30 37.5 32
Sand angular 37 39 36.5
Coal angular 37.5 41 34

The "dynamic" angle listed in this table is determined via an experiment


involving a rotating drum, according to a procedure that will be described in
detail in a subsequent Section.
• The second factor is inherent to the physics of a granular pile. It is best de-
scribed by means of a simple experiment. We use a cylindrical container with
transparent windows at both ends and fill it half way with fine sand-this is
important because a large number of particles is required for the experiment
to work cleanly. We now set the cylinder horizontally and rotate it around its
symmetry axis as shown in Figure 75. At a very low rotation speed (0.01 rota-
tion per second is a good value to try), the free surface of the sand ostensibly
makes an angle e with the horizontal. This angle is what we have called the
embankment angle. Upon closer scrutiny, however, it turns out that this angle
is not unique. It characteristically increases up to a value em, called the angle
of movement, at which time a miniavalanche takes place and e drops back
down from em to a smaller value e,. named angle of repose. The difference
8 = em - er is referred to as the relaxation angle. Its value is typically about
2° for dry granulars.
122 4. Granular Media in a State of Flow

Angle of movement Angle of repose


before relaxation after relaxation

FIGURE 75. Definitions of the angle of movement em and the angle of repose er . The
difference.5 = em - er is called the relaxation angle. Its value is typically about 2°.

This series of avalanches, manifesting more or less periodic relaxations of the


system between a maximum and a minimum value of the embankment angle,
constitutes one of the fundamental properties of the physics of sand piles. As
we shall see shortly, the distributions of sizes and times of occurrence of suc-
cessive avalanche events are fascinating problems in their own right, although
they are beyond the scope of the elementary physics of granulars treated in this
book.
The conclusion emerging from these experiments is that it is a priori impossible
to define unambiguously an embankment angle without a detailed knowledge
of the history of the pile. In particular, it is crucial to know if the pile has just
undergone an avalanche, in which case the embankment angle will be equal to
the angle of repose er • If, on the other hand, the pile is in a critical metastable
state in which the slightest disturbance is likely to trigger an avalanche, then the
embankment angle will be equal to the angle of movement em. Strictly speaking,
the angle of movement is the only one that characterizes the critical state of a
sand pile. These issues are in many ways reminiscent of several observations we
made earlier concerning the indeterminacy of friction forces and hysterisis effects
characteristics of granular flows (see Section 2.3).
It would be desirable to gain a better understanding of the nature of the relaxation
angle which plays such an important role in determining the angle of embankment.
In the absence of any information on how a particular pile was assembled, the
embankment angle is clearly undetermined within a window 8.
Interestingly, Reynolds proposed an explanation for the relaxation angle as early
as 1885 [17]. On the basis of his principle of dilatancy, he argued that particles
in a metastable state (when e = e,) can move about only to the extent that they
can find some free room to do so, which requires that the pile be dilated. This
implies an additional inclination up to the value em' Reynolds in fact calculated
that the relaxation angle 8 corresponded to the extra inclination necessary to dilate
the upper sheets of a pile.
Several considerations can shed light on this important issue. The first has to
do with the behavior of a stack made of a small number of particles. The second
4.1 A Sand Pile in Equilibrium: The Angle of Repose 123

FIGURE 76. In a pile made of a small number of particles, the relaxation angle 8 goes to
zero.

concerns the transition from a regime of intermittent flow to one that is continuous.
We go on to examine these two problems.

The Embankment Angle ofa Pile Made of a Small Number ofParticles


It is not too hard to envision that in a pile made of a small number of spheres, the
difference between angle of repose and angle of movement may be artificially re-
duced, if not completely eliminated. With reference to Figure 76, D is the diameter
of individual spheres and L is the length of the path of an avalanche. If we start at
the angle of repose er , adding a single sphere may be enough to exceed the angle
of movement em. Under such conditions, the two angles become indistinguishable
since both correspond to the onset of an avalanche.
Let us designate by N the number of spheres in a three-dimensional pile assumed
to be in the shape of a cone with an angle close to the critical value. N is related to
the relevant linear dimensions by (L / D) :=:::; (3N) 1/3. Within these approximations,
the relaxation angle (, may not exceed the value (3N) -1/3. This means that there is
a minimum number N min of particles below which the relaxation effects described
above will cease to be relevant for purely geometrical reasons. This number is

r-3~3'
given by

Nmin = ~ (~
We have pointed out earlier that (, is typically equal to 2° for real piles involving
a large number of particles. N min is then approximately equal to 8000 granules.
These considerations show that in matters related to the embankment angle, it is
important to keep track of the number of particles constituting the pile.

Transitionfrom Intermittent to Continuous Regime-Power Laws


When the cylindrical drum described earlier is made to rotate at constant speed Q,
several interesting phenomena turn up [75]:
• For slow rotations (typically less than 0.1 revolution per minute, or rpm),
there is an intermittent flow in which the upper sheet of material oscillates
back and forth between the two angles em and er defined earlier. Varying
quantities of granulars are driven toward the bottom of the pile during each
124 4. Granular Media in a State of Flow

(a) (b)

FIGURE 77. Profile of the free surface of a pile in a slowly rotating drum. (a) The surface
is essentially straight when the rotation speed is just beyond the threshold for continuous
flow. (b) At higher speeds, we witness the appearance of a surge wave, as the influence of
the boundaries at the top and bottom of the flowing layer become noticeable.

avalanche according to a statistics which we will analyze in some detail in


a later part of this chapter. We note that, in this regime, the free surface is
substantially straight, as depicted in Figure n(a), which makes it possible to
define an angle e at any time with reasonably good accuracy.
• As the rotation speed increases (typically to 5 rpm), the flow becomes contin-
uous. Furthermore, the free surface takes on the appearance of an S-shaped
curve illustrated in Figure neb). This phenomenon is somewhat analogous
to the surge wave observed in hydrodynamics. A simple calculation shows
that the particles located near the outer parts of the flowing sheet are thrown
free of the pile by the centrifugal force; they then follow a parabolic trajec-
tory which can become quite noticeable at even higher speeds. As long as
the diameter of the cylinder is sufficiently large compared to the diameter of
the granules and does not rotate excessively fast, an angle e, now called the
dynamic angle, can still be defined fairly accurately.
We now consider what happens when we start from a low-speed regime, when the
flow is intermittent, and gradually increase the rotation speed. A definite transition
to a continuous-flow regime takes place at a rotation speed Q+. If we now slow the
drum back down, the intermittent regime reappears at a speed Q_ that turns out to
be lower than Q+. The two speeds can be substantially different, being about 0.25
and 0.50 rpm, respectively, in the case of dry sand.
This hysteresis effect is depicted in Figure 78. Its origin begins to make sense
as soon as we realize that the time it takes for a granule to fall down is not the

Intermittent Continuous
flux " flux

Q Q+
- - - - - + - 1- - - - + - 1~----> Q

FIGURE 78. The hysteresis effect between intermittent and continuous flow. The arrows
indicate whether the rotation speed of the drum is being increased or decreased.
4.1 A Sand Pile in Equilibrium: The Angle of Repose 125

same in the intermittent and in the continuous regimes, for which the fall times are
designated f) and f2, respectively. The transition from intermittent to continuous
flow occurs when the fall time of a granule becomes synchronized with the interval
T separating two successive avalanches. As we will see shortly, T is actually a
random variable. Nevertheless, its fluctuations are small enough for the argument
to hold. In particular, the regimes switch over when Q+ = 8/f) and Q_ = 8/f2,
with f) < f2.
While the angle of movement 8m is clearly a fundamental property of sand and
evidently reflects a critical phenomenon, things are not so clear when it comes
to the angle of repose 8r after the system has relaxed. Even for large-sized piles,
the angle of repose is generally affected by the fact that the flowing sheet has to
come to a stop at the bottom of the drum. Such finite-size effects are less of a
nuisance than in the case of small-sized piles, but they are real nonetheless. The
influence of the wall must be taken into account if we are to develop an accurate
description of a phenomenon that turns out to be far more complex than it appears
at first.
Since we are dealing with a critical phenomenon, it is logical to want to examine
how the flux of material varies as a function of the inclination angle 8. ) In particular,
the question is whether this particular phenomenon might fit in with other critical
transitions by obeying a law of the form

(4-1)

where J is the flux of granules and 8e is the critical angle.


The flux J of matter is simply given by the law of conservation of transported
matter in steady state, which for a half-filled drum leads to

(4-2)

where L is the length of the drum, and R its radius. Note that J is expressed as a
volume per unit time.
The results shown in Figure 79 correspond to a 19-cm-Iong drum half-filled
with particles of diameter 0.3 mm. Practical considerations impose that the rota-
tion speed Q be restricted to a range between 0.5 and 12 rpm. The lower limit
corresponds to the onset of continuous flow, while at the upper end a substantial
fraction of the flowing sheet is sent flying off due to the centrifugal force. The
experiment consists simply in measuring the inclination angle of the flowing sheet
as a function of rotation speed of the drum. Since we are looking for a power law,
the results are plotted in a log-log format. The graph shows that in the range of
variables investigated there is indeed a power law of the type J ex (8 - 8e n , with r
m = 0.5 ± 0.1. It is interesting to look for the physical significance of the expo-
nent m. Some useful insight can be obtained by reviewing the case, well known in
hydrodynamics, of an ordinary flowing liquid (Brownian flow).

lThe angle e, called dynamic angle, defines one of the characteristics of avalanches. It obviously
depends on the speed of rotation. Under certain conditions, it also depends on the size of the granules.
We will see an application of this poorly understood property in Section 5.4.
126 4. Granular Media in a State of Flow

Sc-S (deg)

20 0
0

0
10

f*
5
Y
4
3
f
2
t
2 3 4 5 10
.n (rpm)
FIGURE 79. Log-log plot of the angular deviation from the critical value versus transport
flux during avalanches in the continuous regime (after [75]).

Power Law for a Newtonian Fluid


Consider a Newtonian fluid of thickness h and viscosity 1] flowing down a plane in-
clined at an angle e with respect to the horizontal. The flux is known to be given by

pgh 3
J = - - sinCe),
31]
where p is the density of the fluid.
In other words, for an ordinary liquid, which happens to have a critical angle
ec = 0, we find an exponent m = 1. The phenomenological explanation for this
result is that the Brownian particles making up the fluid have an instantaneous
velocity far exceeding the net mean speed of the flow. If so, the viscosity 1] reflects
the loss of momentum resulting from multiple collisions. In this context, what can
be said of a granular flow?

Power Law for a Granular Surface


Bagnold suggested that in the case of non-Brownian particles, and unlike ordinary
liquids, energy is dissipated via two different mechanisms [76]:

(1) The rate of collisions between particles is proportional to the velocity gradi-
ent V v. What this means is that the greater the difference in velocity between
adjacent sheets, the more frequent the collisions between particles.
(2) The loss of momentum upon each collision is also proportional to Vv.
In other words, the greater the relative speed between two particles, the
more dissipation upon impact. We might note that this is hardly consistent
with the elementary picture of frontal collisions described in Section 2.2.2.
4.2 Avalanche Models 127

Indeed, it can be argued that a more realistic descliption would involve


nearly tangential collisions between sheets of particles lUshing down a slope
with different velocities.

In any event, according to this model, the friction force resulting from a velocity
gradient V v is proportional to (V v)2. We can thus write the equilibrium between
the friction force and the force of motion as

av)2
-ex ( az + pgz[sin(B) - JLcos(B)] = 0, (4-3)

where JL = tan(Be ) represents the friction coefficient of a granular sheet sliding on


its neighbor, as it was defined by Coulomb. The z-axis is directed downward along
the force of gravity.
Assuming that p and JL are slowly varying functions of the velocity v, we can
write a series expansion of the trigonometric functions in the vicinity of Be up to
first order in (B - Be). After integrating, we get

2
v(z) = '3

This explains the J ex (B -Be) 1(2 dependence suggested by de Gennes and observed
experimentally. Based on this elementary calculation, Bagnold's law in (VV)2 and
a simple law of dry friction are all that is needed to account for the exponent !
characterizing the behavior of a granular flow near the critical point.

4.2. Avalanche Models


Avalanches are perhaps the most extensively studied phenomenon in the physics
of sand piles. As mentioned earlier, systems on the verge of avalanching confront
us with a number of interesting problems that reach beyond the restricted domain
of granular materials and overlap with other branches of physics. 2 Whether it
involves a critical state (described by the type of laws that govern phase changes)
or a simple instability in the sense of mechanics, the behavior of a granular surface
undergoing avalanching continues to be the object of numerous investigations and
spirited debates.
Without going into the details of the controversy, which is far from being re-
solved, we will present three different approaches that are fundamentally different.
Our purpose is not to take sides in favor of one or the other, but simply to illustrate
how different strategies can be brought to bear to explore the same phenomenon.
We will start with a cellular automaton model that makes it possible to predict the
statistical properties of an avalanche. We will follow with a simple macroscopic

2The phrase "system on the verge of avalanching" refers to a situation in which the free surface of a
granular makes an angle e between em and er with respect to the horizontal.
128 4. Granular Media in a State of Flow

ill
FIGURE 80. Principle of the one-dimensional cellular automaton model (CAM) (after
[77]).

model that establishes a relation between the slope of a pile and the flux of parti-
cles in motion. Finally, we will examine a model based on coupled variables that
accounts satisfactorily for a number of characteristics of avalanches.

4.2.1. Cellular Automaton Model (CAM)


The model was originally proposed by Bak, Tang, and Wiesenfeld (BTW) specif-
ically for the purpose of studying systems in a self-organized critical state [77].
Nonetheless, the cellular automaton model (or CAM for short) is directly re-
lated to avalanche processes, although some caution is warranted in making this
generalization. 3

The Principle
Squares are stacked on top of each other to form contiguous columns according
to a set of extremely simple rules:

(l) The height difference between two adjacent columns cannot be greater than
two units. This effectively simulates the angle of repose, which cannot
exceed a critical value without collapse.
(2) When a column rearranges itself because of excessive height relative to its
neighbors, it involves the movement of a set of two unit cells. This is akin
to the domino effect in an avalanche.

Any starting configuration is allowed to relax back to its equilibrium in accor-


dance with the rules stipulated above. The end result is a stable system, such as is
illustrated in Figure 80, which becomes the starting configuration for subsequent
experiments. Individual squares are then added at random to the edifice, one at
a time. Each such event triggers a relaxation process of its own, subject to the

3In this chapter, the term "self-organized criticality" (SOC) applies to a system evolving spontaneously
toward a critical state with no memory of the initial conditions. In other words, it is a system for which
the critical state is an aUraetor as far as its dynamic properties are concerned. The interested reader
is referred to the abundant literature on this topic. There has been much debate about whether this
model is at all pertinent to avalanches. Suffice to say that caution is strongly advised, if for no other
reason than because of the existence of two different angles em and er , rather than a single one.
4.2 Avalanche Models 129

same rules. Since the bottom surface is assumed to be of finite size, squares can-
not accumulate there but, instead, drop off the base. The objective of this thought
experiment is to count the number of squares that drop off after each new square
is added to the pile. The release of one square can trigger mini "avalanches" of 0,
2, 4, and so on, squares escaping off the side. As it turns out, small avalanches are
far more frequent than larger ones.
This elementary CAM process, described here in one dimension, can be gen-
eralized to the case of 2, 3, or more, dimensions by developing suitable computer
algorithms. As artificial as such generalizations may seem in terms of describ-
ing real avalanches, they are of considerable interest because theories can often
predict the value of the exponent in the power laws applicable to any dimension,
even higher than 3. Exercising such algorithms often requires the use of parallel
computing so as to cut down on the computation time and minimize round-off
errors which are the bane of conventional sequential machines.
As an example, we can go through the algorithm describing the sequence of steps
in the one-dimensional case discussed above. The difference in height between
adjacent columns is defined as Zn = hen) - hen + 1). When a single square is
added to the nth column, it entails the following changes:
Zn ---J> Zn +1
and

Whenever the height difference becomes greater than a critical value Zc, the system
relaxes back in such a way that

and
Zn±l ---J> Zn±l - 1
The pile is bounded on the left (corresponding to the index 0), and open on the right
(index N), where squares drop off the pile. Therefore, events at the boundaries are
written as
Zo = 0,
ZN ---J> ZN -1,
ZN-l ---J> ZN-l +1 for Zn > Zc.
It is easy to count the number of stable configurations of a pile made of N columns
by writing the stability condition Zn < Zc, where n = 1,2,3, ... , N. This
gives a total number Z;: of stable configurations. They are not all equally stable,
however. A simple technique to arrive at the least stable of these configurations is to
artificially construct a system in which all columns are unstable (meaning that Zn >
Zc for all values of n), and let it relax spontaneously to some equilibrium which
becomes the fresh starting point from which new squares are then added. Such a
minimum-stability configuration is a critical state in the sense that any subsequent
130 4. Granular Media in a State of Flow

disturbance can propagate through the entire pile. This situation is, therefore,
analogous to the well-known problem of percolation in one dimension. The issue
of stability is considerably more complicated in structures with dimensions greater
than 1. It is decidedly beyond the scope of this book, but we refer the interested
reader to the many articles on this topic published in the wake of the original BTW
paper. Extrapolating the algorithm itself, on the other hand, to more than one
dimension is straightforward. We will do it here for the two-dimensional case. For
a square pile described by Z (x, y), the equivalent of the chain of events considered
previously becomes

Z(x - 1, y) ~ Z(x - 1, y) - 1,

Z(x, y - 1) ~ Z(x, y - 1) - 1,

Z(x, y) ~ Z(x, y) + 2.
If the height difference exceeds a critical value Zc, we have

Z(x, y) ~ Z(x, y) - 4,

Z(x, y ± 1) ~ Z(x, y ± 1) + 1,
Z(x ± 1, y) ~ Z(x ± 1, y) + 1 for Z(x, y) > Zc.

As soon as we leave the comfort of a one-dimensional situation, the connection


with an avalanche of granules becomes less transparent. As noted earlier, however,
extending the algorithm to higher dimensions makes it possible to test the validity
of a model describing self-organized critical phenomena by examining the relevant
power laws.

Implementations ofthe Cellular Automaton Model (CAM)


With sufficiently powerful computers, it is possible to calculate the probability
of occurrence D(s) of an avalanche (the less specific term of "slide" is often
used) of size s over several orders of magnitude of s. It is then possible to plot
lOglO[D(s)] against loglO(S), which immediately reveals any underlying power
law. The "experimental" finding is that

D(s) ex: s-r,

where the exponent T turns out to be approximately equal to 1.0 in two dimensions,
and 1.37 in three dimensions.
It is also of interest to examine the probability distribution function of the
durations (or lifetimes) of these avalanches. In other words, given a system in a
critical state on which an additional square is dropped randomly, how long will the
ensuing slide last? Intuitively, we expect some degree of correlation between the
size of a slide and its lifetime. An event triggering a large number of cells to fall
down the side ought to take longer than another one triggering just a few. Although
it takes a little more work to prove it than in the case of the size, "experiments"
4.2 Avalanche Models 131

show that lifetimes do in fact obey similar power laws, expressed in the form

D(T) ex T-"', (4-4)

where ex is found to be 0.43 and 0.92 in two and three dimensions, respectively.

Lifetime and II! Noise


Without reproducing the calculations in the BTW paper, we will review here
a few simple relations between the fluctuations of a system experiencing small
perturbations and the power law just outlined.
First of all, it is important to understand that the noise spectrum of a system
in a critical state and subjected to various destabilizing disturbances (such as
temperature fluctuations, for instance) contains fundamental information about the
dynamic properties of that critical state. It is intuitively evident that a system with
a large inertia will exhibit low-frequency fluctuations, while a light system will be
able to respond to high-frequency perturbations. As such, the noise characteristics
of a system contain information about its dynamic response.
We have just seen that the durations of avalanches is distributed according to a
power law with an exponent ex. We may therefore also expect a power law in the
reciprocal space of frequencies. It is common usage in electronics to talk about
the power spectrum of the fluctuations, i.e., the probability distribution function
of the square of the amplitude of these fluctuations. The power spectrum is then
written in the form

(4-5)

A simple calculation based on Fourier transforms shows that f3 = 2 - ex. Using


the values of ex indicated previously, the noise power spectrum should involve
exponents of 1.57 in two dimensions, and 1.08 in three.
Before pursuing this analysis, it is useful to recall a few elementary notions
related to various forms of noise commonly encountered in physics [78]. They are
illustrated in Figure 81. White noise has a power spectrum that is independent of
frequency (power law of the type fO). Theoretically, it extends from zero to infinite
frequencies, and there is no correlation between any two frequencies. In reality, of
course, the observable spectrum is inevitably truncated by the finite bandwidth of
the detection system. Typically, the response of the system is flat only between a
lower and an upper limit. 4 Brownian noise, on the other hand, is strongly correlated
and favors low frequencies.
Noise with a 1/f power spectrum shows up in phenomena as diverse as the
occurrence of earthquakes, the blinking of stars, vehicular traffic on highways,
and many others. The apparent universality of 1/f noise has attracted a great deal

4A good-quality hi-fi amplifier, for instance, transmits all frequencies between 20 Hz and 20 KHz
without distortions. We might wonder what the frequency spectrnm of the output noise of such an
amplifier might be. Under ideal conditions (with all other noise sources eliminated), the noise turns
out to have a II! spectrum. This remarkable property is due to the shot noise associated with the
input impedance of the amplifier.
iiiii
132 4. Granular Media in a State of Flow

lIt) S(!J-I/( fJj 2

p'EJwer
Spectrum

-----------
11 fO
Low-correlation shot noise ~

rS229 ~
I~~~~~~~'''~rn Time I log f

FIGURE 81. Three types of commonly encountered noise, and their power spectral densi-
ties (f0, r 1, and r 2) (after [78]).

of attention on the part of researchers. This type of noise is one of the characteristics
of systems that enjoy self-similarity, also known as fractal objects. The reason is
that the noise power in a small frequency window df is given by S(f) df; when
S(f) = Ilf, this noise power becomes df/f. In other words, the noise power
remains invariant provided the bandwidth df is scaled to match the mean frequency
f [79]. That is one of the fundamental properties of so-called self-similar systems,
which exhibit the same properties on all scales.
As we noted in connection with (4-5), the cellular automaton model (CAM)
leads to a frequency probability distribution function very close to a 1If law. This
immediately suggests the possibility that the behavior of such automatons may be
assimilated with that of a self-similar system. As we are about to see, this turns
out to be a complicated issue whose answer is not black and white.

The Statistics ofAvalanches


We have stated in Section 4.1 that real granular piles are characterized by no fewer
than two critical angles em and er as long as the pile is made of a large number
of particles. If the number of particles is small, the two angles merge into one.
We may therefore expect different behaviors depending on the number of particles
involved. This is precisely what is observed in actual experiments.
Several research groups, inspired by the results just described, have tried to ver-
ify that the distributions of real avalanches conform to the power laws suggested
by the cellular automation model. To that end, they have devised a number of
experiments based on different principles, illustrated in Figure 82. In Figure 82(a),
particles are dropped one at a time on a cone-shaped pile supported in the pan
of a balance interfacing with a computer. The weight of the particles falling off
the pile can be measured with reasonable accuracy and gives an estimate of the
4.2 Avalanche Models 133

(a) (b) (c) (d)

-I r -I r f7
Scale Condenser Condenser Microphone

FIGURE 82. Four techniques used to study the statistics of avalanches (after [20]).

size of successive avalanches. In Figure 82(b), a condenser is used to count almost


one-by-one the particles as they roll off a pile that is being resupplied continually.
A condenser is also used in Figure 82(c) to count the number of particles escaping
from a partially closed half-cylinder as it rotates slowly around its axis. Finally,
in Figure 82(d), a drum almost half-filled with a granular material spins slowly
around its axis, causing periodic avalanches of varying sizes. A tiny microphone
placed near the apparatus records the noise of these avalanches. All these meth-
ods have yielded results that clearly depended on the size of the population of
granules.

Piles Involving Large Numbers ofParticles


The experimental results shown in Figure 83(a) indicate that, with a large popula-
tion of particles, avalanches occur according to a nearly periodic pattern. The data
were collected with an apparatus of the type depicted in Figure 82(c) [21], [80].
Microphone-based experiments further reveal that the distribution of successive
avalanches is far from being as important a parameter as suggested by SOC models
[66]. The reader may recall that the model predicts many more small avalanches
than larger ones. The statistics actually observed in experiments deviates substan-
tially from the predictions of the model.
The point is clearly illustrated in Figure 83(b), which displays a bell-shaped
distribution function reminiscent of a Gaussian statistic, quite different from the
dotted straight line expected on the basis ofthe theory. The experimentally obseved
distributions of avalanche sizes and durations appear more consistent with first-
order transitions, rather than the second-order transitions predicted by models of
self-organized critical systems.
Yet, on closer examination, two phenomena observed in real experiments may
well explain, at least in part, these discrepancies:

(1) Many avalanches, particularly small ones, never make it to the bottom of the
slope and, as such, never get recorded. The reason is that matter accumulates
at the bottom of the structure, which tends to diminish the angle the free
134 4. Granular Media in a State of Flow

~;~J 0 1 2
t (ks)
2.------.----r----,.-------,

{j) 0
C;
..Q
-2
(b)
-4
-3 -2 -1 0
lag(r)

FIGURE 83. Statistics of avalanches observed in a rotating drum. Diagram (a) shows the
number of particles transported as a function of time as the drum rotates at a constant speed
of 1.3 degree/min. Diagram (b) shows the power spectral density. The dotted line indicates
a 1/f dependence. The granular material is made of particles 0.5 mm in diameter (after
[80]).

surface makes with the horizontal, thereby dropping it below the critical
value, with drastic repercussions for the physics of the process. This is a
classic example of a finite-size effect. It acts somewhat like a high-pass
filter, as it were, favoring larger events.
(2) As noted above, the splitting of critical angles into two values-angle of
movement and angle of repose-implies that a system that has just experi-
enced an avalanche is no longer in a critical state. It is necessary to incline
it by an additional angle em - er in order to reach a new critical state. It is
then not obvious that real avalanches involving many particles truly con-
stitute critical systems, even if they can still be considered self-organized.
In light of some of our previous observations, however, it seem likely that
small avalanches may not suffer from these restrictions. To some extent, the
experiments we are about to describe support this view.

Piles Involving Small Numbers ofParticles


As we learned in Section 4.1, a pile made of a small number of granules (typically
less than 8000) cannot-for purely geometrical reasons-exhibit two distinct em-
bankment angles em and e,.. Under these circumstances, it may be permitted to
expect that the behavior of such a system should conform more closely to the clas-
sic cellular automaton model described in Section 4.2.1. Experiments conducted
with a few hundred particles in an arrangement of the type of Figure 82(a) effec-
tively confirm this [78]. The results are reproduced in Figure 84. The experiment
is intended to reproduce as faithfully as possible the conditions of the cellular au-
tomaton model. Granules are dropped, one by one and randomly, on a cone-shaped
4.2 Avalanche Models 135

10 100
Avalanche size (number of particles)

FIGURE 84. Results of experiments on limited-size piles made of different particles. The
fractional rate of avalanche occurrence (i.e., the number of avalanches of size s normalized
to the total number) is plotted vertically against the number of particles s involved in the
corresponding event. In (a), the triangles correspond to steel balls, and the circles to glass
beads. In (b) the squares refer to polystyrene beads and the circles to glass beads (after
[78]).

pile supported in the pan of a balance. The balance keeps track of the weight of
material falling off the pan during each avalanche event. We may expect that the
shape of the pile should playa fairly important role, since a convex pile is inherently
two-dimensional, in accordance with the model developed in Section 4.2.1.
In a configuration of this type, the sizes of successive avalanches do exhibit a
considerable dispersion [78], [81]. Small-size avalanches are plentiful, while large
ones are relatively rare. Figure 84 also reveals that, regardless of the nature of
the materials (steel, glass, or polystyrene), the statistics of avalanche sizes indeed
conforms to a power law, as predicted by the cellular automaton model. It may be
objected, however, that the agreement has been demonstrated over not much more
than one decade, which some may deem too small a dynamic range to truly test
the validity of the law.
Although several experiments conducted under different conditions have all
confirmed that the behavior of a granular pile differs according to whether it is
made of a large or a small number of particles, the reasons for the crossover
from one regime to another are still not fully elucidated. It is possible that a
more sophisticated cellular automaton model may be able to account for such
finite-size effects. Yet, the existence of the two angles em and en as well as the
relaxation oscillations between these two extreme states, constitute a fundamental
objection difficult to overcome when modeling a dry granular material. A number
136 4. Granular Media in a State of Flow

of researchers have tried to artificially induce a relaxation of the embankment


angle at frequencies higher than that of the natural relaxation described earlier.
This approach is presented next. It makes it possible to introduce a notion that
remains somewhat controversial, namely, that of granular temperature.
Relaxation ofthe Critical Angle-Granular Temperature
We now describe an experiment based on the fact that a pile inclined at an angle em
is inherently unstable [21], [80]. As we have already emphasized, em characterizes
the truly critical state of the system, while er corresponds to a pile in a relaxation
state. The idea behind the experiment is to subject a pile on the verge of avalanching
to a vertical vibration sufficiently strong to promote an artificial relaxation. The
intent is to let the system stabilize in a single state that may be thought of as
supercritical. Looked at in a different way, the vibration is meant to artificially
reduce the difference 8 = em - er •
Consider a pile containing a large number of particles placed in a cylindrical
drum, as depicted in Figure 75. The drum is placed on a base that is set in vibration
by a loudspeaker excited by a sinusoidal electrical current. The concept is illustrated
schematically in Figure 85(a). In a first experiment, the rotation speed Q is low
enough for the intermittent regime to prevail, in which case the pile is inclined
by an average angle ass with respect to the horizontal (the subscript ss stands
for "steady state"). As expected, the value of ess decreases as the amplitude of the
vibration increases. The angle ess effectively constitutes a measure of the sinusoidal
excitation and becomes a parameter for the experiment.
Keeping the rotation speed low (1.3 degree/min, which ensures an intermittent
regime), we can monitor the size distribution of avalanches just as we did earlier

(a) (b) (c)

.
............................. 24

21
.... I/A ..... .o\l,
~..
....
~ ..
'2 "........
o _, 18
.... .....
1
~
g; 15

~ 12
ff
Vibrations -3 -2 .. , 0
.9u =26.1°
.. 8..=22.2°
loglOCf(sec- 1)J
.9..=19.5°
.9..=16.4°

FIGURE 85. Results of experiments done with a rotating drum subjected to vertical vi-
brations. The angle ess, which corresponds to the steady-state angle obtained for a rotation
speed of 1.3 degree/min, is a measure of the excitation intensity. The dotted line in (b)
indicates the 1/f behavior predicted theoretically. In (c), Q = 0, and the structure relaxes
according to a log(t) behavior, whereas the CAM model predicts a relaxation proportional
to t (after [80]).
4.2 Avalanche Models 137

in the static case. The question we seek to answer is whether vertical vibrations
can put a pile in a supercritical state characterized by a single critical angle, such
that the power law predicted by the cellular automaton model is observed. Based
on the curves shown in Figure 85(b), the answer is decidedly no. Even when the
avalanche angle, which started out at 39°, is lowered by more than 2°-enough
to cancel the difference 8-the power spectrum of the avalanches continues to
exhibit a peak, which is inconsistent with the SOC model. As the critical angle is
lowered by even greater amounts, the power spectrum begins to look more like a
power law, but with an exponent that deviates substantially from the 1If behavior
expected theoretically. The experiments suggest a dependence closer to II fo. 8 .
Nevertheless, this is not enough to conclude that the poor agreement between the
CAM model and the real systems is solely due to the existence of the two angles
em and er •5
Other measurements done with the same apparatus provide a wealth of infor-
mation. The pertinent results are shown in Figure 85(c). Here the drum no longer
rotates, but it is stopped in a position such that the free surface of the granular
material is at an angle e just below the value er • When the vertical vibration is
then turned on, the angle of repose turns out to decrease (or relax) with time.
Furthermore, the rate of relaxation depends on the amplitude of the current in the
loudspeaker. The results shown in Figure 85(c) constitute fairly compelling evi-
dence that the embankment angle e evolves as log(t), while the CAM-SOC model
suggests that it should vary proportionally to t. This rather remarkable result has
been interpreted in terms of a simple model involving a relaxation process induced
by an agitation (caused by the loudspeaker) that may be considered the equivalent
of a "thermal" agitation.
In the context of this model, the intensity of the vibration plays the role of an
"effective" temperature Teff [21], [80]. There is a perhaps rather bold analogy to be
made with the phenomenon of electrical conductivity, where a flux j of electrons
results from an applied electrical field E through the bias of the conductivity a,
the latter being of a magnitude governed by the degree to which electrons are
released from randomly distributed traps. Here, the electric field E-the driving
force-is assimilated with the angle e, and the current density j with the rate of
change deldt. By analogy with the electrical case, a granule can be thought of
as trapped, much like electrons in a conductor, by neighboring granules. Rather
than reviewing the many assumptions involved in this problem, we will simply
calculate the average height of an effective barrier V as a function of the angle e.
Expanding V to first order about the starting point of the experiment, we may write
V ~ V o + VI (e - er ). Knowing that a flow is triggered when e = em, we impose
the condition V (em) = 0, which leads to 8 == e - er = Vol VI. Pursuing our
electrical analogy, we write that the rate of change del dt, like the current density,

5 As we will show in Section 4.2.2, the dragging effect of the rotating disk, even as slow as it is in
this case, can introduce an artificial periodicity in the sequence of avalanches. This point is still
being actively debated. Nevertheless, very recent experiments have for the most part confirmed the
conclusions presented in this paragraph.
138 4. Granular Media in a State of Flow

depends exponentially on the energy barrier V as V I kTeff

de
- = -Ae exp[tJ(e - er )],
dt
where both A == A o exp( -Vol kTeff) and 1'3 = VI! kTeff are independent of e.
The solution of the differential equation can be expressed by means of the
exponential integral function E 1 (tJe). Approximations valid when the argument
M» I lead to
I
e ~ er - -loglQ(tJAert + 1),
1'3

which is consistent with the loglQCt) behavior observed in Figure 85(c) for times
greater than to = 111'3 Aer . The agreement remains reasonably good even for shorter
times. This derivation thus accounts for the loglQCt) behavior of the embankment
angle of a vibrated granular pile.
Interpreting avalanches in terms of thermal detrapping is fundamentally different
from the critical angle model proposed by BTW. The experiments described above
clearly show that the simplistic view of the angle of repose er as a critical angle
does not stand the test of a detailed analysis of the relaxation mechanism.

4.2.2. Stick-Slip Model ofAvalanches


Chronologically, the model we are about to present came into existence after the
ones described above (see Fauve et al. in [27]), against a backdrop of mounting
evidence that the SOC model had serious shortcomings. It starts from an altogether
different point of view. The underlying idea is remarkable simple. It is based on
a system of phenomenological coupled equations describing the behavior of two
pertinent observable quantities, namely, the angle e of a sheet in motion and the
flux D of particles sliding down a slope during an avalanche. 6 The basic principle
is embodied in a particular relation between friction force and velocity and in
the coupling of friction with a spring, as outlined in Section 3.1.1. As such, it
takes advantage of an explicit parallel, already suggested in Chapter 3, between
the alternate stick-slip motion related to solid friction and intermittent series of
avalanches.
Figure 86(b) illustrates the approach. It shows a pad rubbing against a surface
with static and dynamic coefficients of friction Ihs and Ihd (x), respectively. As
usual, Ihd is smaller than Ihs and can depend explicitly on the velocity, as explained
in Section 2.3. The pad is also connected to a spring of stiffness K and moving
at a velocity V. This arrangement is entirely similar to that in Section 2.2.1, and
was subsequently used to analyze the phenomenon of hysterisis in Section 3.1.1.
If i; designates the deformation of the spring and Q) the speed of rotation of the

6The paTticles in flow are sometimes referred to as a "moving species." The idea of describing the
problem by means of these two coupled variables has recently been exploited from a different angle.
It has led to several interesting developments which we will discuss later on.
4.2 Avalanche Models 139

(a) (b)
FIGURE 86. Illustration of the correspondence between (a) avalanche processes and (b) the
stick-slip mechanism. Both phenomena can be described by the same set of differential
equations (after [27]).

cylinder depicted in Figure 86(a), it turns out that the two systems are governed
by the same set of equations, with the following correspondence

w +------+ V.

In particular, when the pad is in motion, the elongation of the spring is given by
the equation

which simply describes the motion of a harmonic oscillator at angular frequency


W = J K / m around its equilibrium position defined by K S = mg lid. As it
happens, though, this analysis is a bit too simplistic when it comes to an interaction
like dry friction. As we learned in Section 3.1.1, the force of traction associated with
the spring at times exactly offsets static friction, at which point the pad suddenly
stops. It then remains "stuck" on its support until the force increases sufficiently
to set it in motion once again. That is exactly the principle behind the stick-slip
phenomenon discussed previously. The condition for sticking can be written as

dx
if -=0 and
dt
Figure 87 illustrates graphically the behavior of any stick-slip mechanism. Upon
stretching the spring at a constant rate, which is equivalent to moving toward the
right on the diagram, the pad is made to execute a periodic motion characteristic
of a relaxation oscillation. On the one hand, the system returns to its equilibrium
140 4. Granular Media in a State of Flow

0=0

·x
o
o
til
OJ
:0 (jJ/y
co
~

<P f <Pd <Ps


Variables e or 1;
FIGURE 87. Visualization in phase-space of the behavior of an interrupted oscillator (after
[27]).

position with its own internal time constant, Ti = 2][.j m/ K, characteristic of


a gliding motion. On the other hand, it is set in motion with a time constant
Te = 2mg(lts - Itd)/ KV that depends on the external conditions. We now return
to the avalanche aspect of the model. We seek the equations that relate the angle
e to the flux D of particles in the mid-plane of the cylinder by taking advantage
of the analogy with the stick-slip pad. When the bed of particles is at rest, we
obviously have
de
-=0 (4-6)
dt '
dD
-=0 (4-7)
dt '
if

D =0 and tan(e) < Its. (4-8)

When a flow occurs, the preceding analogy allows us to use the equations estab-
lished in Sections 2.2.1 and 3.1.1
de
-=w-yD,
dt
dD
dt = P[sin(e) - Itd(D) cos(e)],

if

D=O and tan(e) > It"

where P = gh, h is the thickness of the moving sheet, and y depends on the
geometry of the flow; to a first approximation, y can be considered a constant. The
first of these equations expresses the conservation of the mass of material driven
4.2 Avalanche Models 141

to the bottom and replenished at the top by the rotation of the cylinder. The second
equation is obtained by analogy with the results found in Section 3.1.1.
The notation can be simplified by introducing an angle cD d defined by

or

We should bear in mind that fJ--d depends on the flux D, which is itself given by

D = _w_-_d_e/_d_t .
y

Eliminating D between the first two of our set of equations yields a second-order
differential equation in e

For small deviations around cDd, we simply have

where p is defined as p = p / cos (cD d). This last equation describes oscillations
of the angle e with a period (l/y p )1/2. The steady-state solution, corresponding
to a constant flow of the rolling species, is given by

and

As expected, the stability of this solution and, more generally, the overall behavior
of the system, is critically dependent on how the coefficient of dynamic friction
fJ--d depends on the flux D.
Expanding cDd(D) to first order around the steady-state value DO, we get

cDd - d)
cD °+ (acD (D - D ) - d) --,
° = cD °- (acD 1 de
aD °y dt
~
aD 0

which can be substituted back into the second-order differential equation to give

(4-9)

As we might have suspected, the behavior of avalanches depends on the slope of


ed (D) in the vicinity of the steady-state value DO. Let us define a parameter a by
a == - ~ (a cDd/a D)o. One of three situations can arise:
142 4. Granular Media in a State of Flow

• Case a = O. The damping term in the differentia1"equation vanishes. We are


then in a stable regime in which the behavior depends essentially on the initial
conditions (stick-slip, continuous flow, or oscillations).
• Case a < O. The oscillations are dampened and a stable state is reached after
a certain number of oscillations have died out.
• Case a > O. The oscillator is akin to a negative resistance and acts as an
amplifier. The solution is unstable and corresponds to a stick-slip situation.

We proceed to examine in more detail the different possible regimes determined


by the functional dependence of the dynamic coefficient fid on the flux D of the
rolling species. Once again, the reader is reminded that fid specifies the value of
the angle <Pd.
When fid-and, hence <Pd-is independent of D, (4-9) then describes a simple
harmonic oscillator whose angular motion is symmetrical about <P d. An avalanche
is initiated at the angle <Ps given by <Ps = tan -1 (fis), and stops at a final angle <P f
that is independent of the rotation speed w. The angle is such that

It is important not to confuse the quantity


,6. <P, which measures angular excursions
during the static phase of an avalanche, with the maximum deviation ,6.(} as it is
actually observed during avalanches. Note that if <Pd = <P s , meaning that there is
no difference between the static and dynamic friction coefficients, the amplitude
of avalanches approaches zero as the rotation speed w goes to zero. By contrast,
if <Pd < <P" avalanches retain a finite amplitude when w -+ 0; it is given by

In plain language, at very low rotation speeds, the relative velocity imparted by the
cylinder becomes negligible, and the angular windows ,6.(} and ,6. <P both converge
toward the difference em - er , which is what we called the relaxation angle in
Section 4.1 and is defined unambiguously only when the rotation speed is slow.
Next, we determine the effect of the rotation of the disk on the duration r of an
avalanche. That quantity can be derived from (4-9) and is given by

r = ffi
2 { rr - tan -1[ffi
~(<Ps - <Pd) ]} .

Here again, we should appreciate that the measured duration 8 T of an avalanche


is not the same as r because of the rotation of the cylinder. As a matter of fact,
8T is exactly half an oscillation period. It therefore corresponds to the relaxation
frequency of the system, which is equal to rr / ffi regardless of the rotation speed.
Not surprisingly, 8T and r become indistinguishable for infinitely slow rotations.
Considering now the case when the dynamic friction coefficient does depend
on the flux D, (4-9) indicates that the D versus () closed trajectories depicted
in Figure 87 become asymmetrical about <Pd. Avalanches stop at an angle that is
smaller or larger depending on whether the coefficient of friction is a decreasing or
4.2 Avalanche Models 143

o
~
a
1:5
CD
.~

I-

o cD d cDs
Angle e
FIGURE 88. Numerical calculations of the trajectories D(e) with a very low rotation
speed (w = 10- 3 rad/s). The coefficient of dynamic friction decreases with speed, which
introduces an asymmetry in the curves, while the angular difference between trajectories
gradually increases (after [27]).

increasing function of the flux D. For instance, in the limit of a small and constant
variation a = - ~ (B<Pd/B D)o, (4-9) can be solved in closed from, giving the result

More general cases can only be solved numerically. Figure 88 shows an example
of such a calculation. Experiments done with noninvasive observation techniques
reveal remarkably good agreement with the present model, in spite of its great
simplicity.7 Further details on the comparison between theory and experiment can
be found in the original article on this topic [27]. By way of example, Figures 89 and
90 reproduce the results of two experiments conducted in the same drum rotating
at two different speeds. The similarities with Figures 87 and 88 are striking.
As a general comment, we might point out that avalanches are rather chaotic
for very slow rotations and tend to become more regularly periodic as the rotation
speeds up. This is suggested by the model and certainly appears confirmed ex-
perimentally. An important consequence of this observation is that it is generally
difficult to carry out reliable and controlled measurements in a rotating cylinder.
The same comment actually holds true for any other configuration. Gradually in-
creasing the inclination angle so as to drive the system out of equilibrium must
always be done as slowly as possible.
As pointed out previously, this particular stick-slip model for avalanches leads
to a second-order differential equation, in which the dissipative term (in de / dt)
involves a nonlinear dependence of the friction coefficient f.-id on the flux D. The

7The term "noninvasive" technique means that the measurements do not disturb the object being
measured. These techniques are most often optical in nature. They include the imaging methods
described in Chapter 3.
144 4. Granular Media in a State of Flow

3
'CD
I 2

o
-1

0.4 0.8 1.2 1.6 2.0


8 (arbitrary units)

FIGURE 89. Optical measurements of the slope of piles and its time derivative for a series
of avalanches. The relation between the amplitude of avalanches and the angle at which
they begin is evident. The trajectories exhibit the asymmetry predicted by calculations. The
rotation speed used here is a rather slow 0.023 rpm (after [27]).

• CD
I

-2
0.5 1.0 1.5 2.0
8 (Arbitrary units)

FIGURE 90. Same as Figure 89, but with a faster rotation speed of 0.52 rpm. The trajectories
become more symmetrical and nearly periodic (after [27]).
4.2 Avalanche Models 145

dependence in question can take on varied fonns, which determine to a great extent
the precise behavior of the flow.
Put another way, the key to any analysis of stick-slip mechanisms is to learn
something about the F (v) dependence, F being the force of friction and v the
velocity. The functional shape of F (v) has in fact been the focus of a great many
investigations, notably among geophysicists and geologists, who are interested
primarily in earthquakes. A comprehensive review of these largely heuristic models
would be beyond the scope of this book. It nonetheless serves a useful purpose to
review a few general principles, which we now proceed to do.
Different Friction Models
Burridge-Knopoff (BK) Pads
The Burridge-Knopoff pad model is in many ways the paradigm among a slew
of approaches aimed at deriving functional laws for F (v). The model does suffer
from some weaknesses, which have inspired numerous variations attempting to
remedy them. We will present only a brief description.
The principle behind the BK model is illustrated in Figure 91 [82]. A series of
pads, each of mass m, rests on a horizontal base. The pads are connected to a rope
by meanS of identical springs of stiffness k. The pads are also interconnected by
a different set of springs with stiffness K. As it moves horizontally at a constant
velocity v, the rope drags along with it all the pads and causes them to slide against
the base (it is important that the pads not bounce off the base during their motion so
as to maintain friction at all times). We sense intuitively that such a system can be
the seat of complex oscillations. Our objective is to find the applicable fonn of the
force of friction, which we know must depend on both k and K. The differential
equation governing the motion of the j th pad, flanked On either side by pads j - 1
and j + 1, reads
2
d_X,
_ J = K(X+ 1 - 2X, + X J'-I) - k(X - vt) - F (dX
_J
,)
dt2 J J J dt '
which leads to an elementary solution of the type
1
Xj = vt - kF(v).

Constant velocity

\ Friction Force F(dxldt)

FIGURE 91. Generic one-dimensional Burridge-Knopoff (BK) pad model.


146 4. Granular Media in a State of Flow

Velocity

FIGURE 92. Dependence of the dynamical friction force on velocity in the BK model.

A detailed analysis of these equations shows that the F (v) has a functional depen-
dence such as is shown in Figure 92. Except very near the origin, the curve F(v)
features a negative slope which, as described earlier as well as in Section 2.3, is
likely to give rise to amplification and possibly chaotic oscillations. This partic-
ular function, and others of similar shape, translate a very common phenomenon
known in geophysics as the velocity weakening friction law. In plain language, it
simply means that the friction forces decrease as the velocity of the relative motion
increases. 8
Other Friction Laws F(v)
Figure 93 illustrates several functional dependencies of F (v) proposed by a number
of authors on the basis of different arguments [48], [83]. Various phenomena such
as contact wear-out and fatigue, frictional heating, and others, can come into play
and give rise to exotic shapes for the F (v) curve. There exists a fairly abundant
literature on this topic, which the reader may want to consult.

A Few Remarks Concerning the Stick-Slip Model ofAvalanches


First of all, the stick-slip model-based as it is on the physics of continuous
media-breaks down when the number of particles involved is too small. The
problem is compounded by the fact that the model makes an explicit distinction
between the angle of repose er and the angle of movement em. As we have seen,
the two angles become indistinguishable with fewer than a few thousand granules.
This relatively elementary model takes into account some properties of real
avalanches by means of the following features:

• The discontinuity of the force of friction when the flux of particles is zero, in
accordance with Coulomb's law of dry friction, which specifies that there are
two different friction coefficients, depending on whether there is movement
or not.
• The existence of a "negative resistance" coupling affecting the function F (v).

8A deadly manifestation of this trend is the phenomenon of aquaplaning of vehicle tries on wet roads.
4.2 Avalanche Models 147

A sphere on an
inclined sheet. V-shaped curve
"Half- W" curve
C C
IC719~)
~
Cal
2:Q O'u
c al

.g1l5 t;:E
'C; ~
LLo LL
(.J U
Velocity Velocity

vv
Contact fatigue:
"Inverted-V" curve N-shaped curve
(Carlson and Langer, 1989) (Barenblatt et al. 1981)

C C

~
c al cal
0'(3 0'(3
"u ~~
"i::
UtE
"C ~
LL LL
U U
Velocity Velocity

FIGURE 93. Various proposed functional dependencies of the friction force on velocity.

• A spatial coupling that accounts for the propagation of the rolling species after
the movement has been initiated. It also explains the duration of avalanches.

Obviously, the model has limitations. For one thing, it does not enable us to
calculate the profile of the granular flux. There exist other, more sophisticated,
models that can do that. They too start from a system of coupled differential
equations, but they involve a different set of variables (in this case, the flux of
rolling species and the height of the pile) and are based on a rather different
principle. We will take the time to briefly discuss the broad outlines of these
models, which are still in the process of being developed. We deliberately limit
ourselves to a phenomenological description.

4.2.3. Avalanche Models Based on Coupled Variables


The basic idea behind such models is to work with two carefully chosen variables,
inspired by those used in deriving the fundamental equations of hydrodynamics
[84]-[86]. We consider the simplest possible case, which is of dimension I + 1.
A single spatial variable x suffices to describe the dimension along which the
flow of particles proceeds. The height h(x, t) of the pile depends on both position
x and time t. At a point of abscissa x, the slope of the pile is measured by the
derivative with respect to x. For simplicity of notation, we denote this derivative
-axh, the minus sign being there to remind us that the free surface slopes down
in the direction of increasing x. A critical situation will be reached when the slope
-axh exceeds a certain threshold Sc that can obviously be identified with the angle
of repose er discussed earlier.
We introduce a second variable R (x, t) corresponding to the density of rolling
species. It characterizes the flux of matter sliding down the slope. It too depends
on both position and time. Some authors have proposed that the density R(x, t)
148 4. Granular Media in a State of Flow

should be governed by a diffusion-convection equation of the form [84], [85]

3R(x, t)
--- = -3x (vR) + 3x (D3 x R) + feR, h), (4-10)
3t
where v represents the velocity of the sheet toward the bottom of the pile, and D is
a diffusion constant that can drive particles in both directions. The first two terms
on the right-hand side of (4-10) are quite familiar; they correspond to the usual
convection and diffusion mechanisms. The last term feR, h), on the other hand,
is the crux of the matter, because it must be able to account for the properties of
avalanches. As such, what form to chose for it bears some discussion. The function
f (R, h) must act as a mathematical operator capable of stopping a particle in mo-
tion or, conversely, setting one that is standing still in motion. Taking our cue from
the phenomenology discussed earlier in this chapter, we may begin to have a feeling
for what this function might look like by imposing a certain number of conditions:

• A granule at rest can be set in motion only if it is dislodged by another granule


already in motion.
• This dislodging process is effective only if the local slope -axh exceeds Sc
(in the context of our previous discussion, this means 8r < 8 < 8m ). In order
to further simplify the notation, we subtract Sc throughout from the slope.
Under these circumstances, the gradient of hex, t) will be a small quantity
at all points along the x-axis. Note that, with our sign convention, axh >
indicates a slope that is less than the critical value.
°
• If 3x h > 0, the granules are no longer in unstable equilibrium, since we are in
a subcritical situation. The dislodged particles then roll down independently
of each other without triggering a large-scale avalanche. Steady state demands
that they come to rest and "stick" to the surface at the same rate R as they are
being supplied.
• If 3x h = 0, that is to say, if the slope matches precisely the critical angle,
first-order derivatives vanish and it becomes necessary to push the analysis to
a;
second-order by including h. We may anticipate that the system, controlled
through the bias of the operator f, will evolve so as to erode any roughness of
the free surface. Put another way, the local curvature will tend to be reduced,
here again in proportion to R.

Given these constraints, all based on what we already know about avalanches, we
are now in a position to look for a plausible form for f. The simplest possible way
for f to comply with our four requirements is to be of the form

(4-11)

where y and K are two positive constants. We might point out that this expression
depends on hand R to first order, which is a distinct advantage from the standpoint
of analytical calculations, as we shall see shortly. We also note that the operator f
is now proportional to R. That is a major difference from the equations considered
earlier. Previously, the flux of rolling species experienced amplification not as a
result of a domino effect, as is the case here, but solely because of the negative
4.2 Avalanche Models 149

curvature of the function a<t>lax in the vicinity of its equilibrium value [see (4-9)].
In the previous stick-slip model, the driving force determining the flux was an an-
gular deviation. Here, the flux is determined both by the deviation from the critical
slope and by the quantity of material already in motion.
In order to account for material at rest, all we need to do is write h = - r in the
previous equation. Thus, the height of the pile obeys the equation

(4-12)

so that the total quantity of matter h + R is conserved locally. We further note that
(4-11) does reproduce the metastability of a pile on the verge of avalanching which,
as we know, is a defining characteristic of granular piles. We can indeed convince
ourselves that, in the absence of any flow, the equation does not spontaneously
generate any avalanche. Thus the surface appears as if frozen in a static equilibrium.
If, on the other hand, we start in such a situation and introduce a perturbation on
the surface of the pile, at least a few grains will start rolling down the slope. The
slide will take place in a finite amount of time and lead to a stable state. That too
can be recognized as one of the fundamental characteristics of avalanches.
It is not essential here to examine further the details of this formalism. Suffice
it to say that it can be made use of either analytically (in the simplest cases) or
numerically (in the tougher ones). The interested reader is encouraged to consult the
basic literature about the theory [84]-[87]. We will restrict ourselves to discussing
two applications of this model.

Upward Propagation of Perturbations


Consider a situation where the free surface of a pile is a straight line inclined at
the critical angle Sc, with continuous and constant flux R o. The model indicates
that a small mound created near the bottom of the slope travels back up with a
constant velocity Vh = Y R o. During the propagation toward the top, the mound
gets broadened and attenuated (because of the presence of the diffusive term in the
equation), as depicted in Figure 94.
It is not hard to envision how such a phenomenon should come about when a
small depression is created near the bottom of the pile. In this case, the upper portion
of the depression acts as a source of miniavalanches, causing a net transfer of matter
toward the bottom, and leaving behind a hole that appears to move upward. The

time t time t+dt

FIGURE 94. The model predicts that a perturbation created anywhere on the slope works
its way back up. In the process, it gets attenuated and widened because of diffusion effects.
150 4. Granular Media in a State of Flow

efficiency of this process as measured by ah / at, and hence its propagation velocity,
depends linearly on y and R o, as is evident in (4-12). Things are far less intuitively
obvious, on the other hand, in the case of a mound. We would be more inclined
to think that the lower part of the mound would collapse. Instead, because it is
based on a set of coupled equations, the model suggests that the material located
above the mound can somehow "sense" the perturbation below and undergo local
minislides that rearrange material in such a way as to generate a new mound that
appears to propagate toward the top. Perhaps this unexpected property is a generic
signature of this type of model.

Simulation ofAvalanches
The ultimate test of the model is, of course, whether it can correctly simulate an
avalanche. To resolve that question, we consider a pile in a metastable state such
that axh(x, 0) = -SoC <0). That can be easily accomplished by increasing-at least
mentally-the inclination of the pile, initially assumed to be at the angle of repose,
until it approaches the angle of movement. Barring any additional perturbation, the
new state is metastable, and nothing happens (as long as the angle of movement
is not exceeded). In this situation, the upper part of the pile is at an angle greater
than the angle of repose.
What happens now if we add a few grains near the bottom of the pile? Those who
have correctly grasped the essence of the correlations associated with the coupled
equations will readily guess that the result is going to depend quite critically on
the value chosen for So.
To clarify this point, we need to consider in some detail how the process un-
folds. Assume that at time t = 0 a localized perturbation is created at point x in
the form of a few moving grains, which may be written as R (x', 0) = flo (x' - x).
Two opposite effects combine to produce the final outcome. First, as we have just
seen, the small perturbation propagates up the slope toward the top of the pile.
That is directly implied by (4-10) in which we neglect the coupling term r. The
ascending perturbation decreases with time as R(x, t) ex exp( -v 2 t /4D), as is
characteristic of all diffusion processes. On the other hand-and that is the second
competing phenomenon-the grains rolling down at abscissa x will dislodge other
grains initially at rest. Equation (4-10) shows that, when v = D = 0 (velocity and
diffusion are both zero), the density of the rolling species grows exponentially as
R(x, t) ex exp(y Sot). Stated in words, the initial perturbation sets off two compet-
ing phenomena that both grow exponentially. The situation is critical in the sense
that either one can win out. We find that if So > Sd ~ v 2 / y D, the number of grains
that are dislodged grows faster than the number of grains that make it back up the
slope. A runaway event-runaway because governed by an exponential-is trig-
gered in the form of an avalanche. 9 In the other case (corresponding to So < Sd),
the diffusive term dominates; it attenuates the effect of the perturbation and the
system remains at rest.

9We have chosen to preserve the notation used by the authors of this model. This should help those
inspired to read the original articles. The correspondence with the angles em and e,., used earlier in
this chapter, as well as their difference 8 = em - e,., is obvious. Here Sc' (== e,.') is set equal to 0, and
Sd(== em) becomes equivalent to what we called the angle of movement.
4.2 Avalanche Models 151

Starting from a simple consideration of an angle of repose Sc', the present model
leads quite naturally to the definition of a critical angle Sd which corresponds to our
familiar angle of movement em. It is the angle required to trigger an avalanche. 10
We might add that numerical solutions of the applicable equations confirm these
interpretations.
The model can, of course, be applied to a variety of situations, from filling
a silo to far more complex scenarios. We must keep in mind, however, that the
formalism accords a great deal of importance to what we have referred to as a
diffusive process, which in actuality boils down to a mechanism of attenuation and
upward propagation competing with the rate at which grains are knocked loose.
De Gennes has pointed out that such a diffusive term was unsatisfactory on at least
two grounds [86]:

• The first has to do with the range of this diffusion. For any variation of the
density of a rolling species R of size L, the diffusive term corresponds to a
perturbation of the order of D / lv, which is itself of the order of d/ L (where d
is the mean diameter of the grains) relative to the convection term. In practice,
this means that the diffusive term must be rather small for small-size slides.
• The second objection is that the physical interpretation of the diffusive term
is a sort of Brownian motion that enables some of the particles to make their
way back up the slope. The problem is that such an interpretation is plausible
on a very small spatial scale only, of the order of D / v ex d, in which case a
continuous description looses its meaning.

This opens to door to the possibility of modifying the model, while still preserving
the spirit of the initial equations.

De Gennes's Modified Model: Rotating Drum Experiment


The previous model suggests that rolling species respond to a small disturbance by
convection, amplification, and diffusion, as long as the slope is inclined by more
than the angle of repose. The magnitude R of the flow at a given location x is given
at some subsequent time t by an equation of the form

Rex exp[y(e - e r) - :~}.


The basic idea behind the modified approach is to keep the general form of (4-10)
and (4-12) but discard the diffusive term. The objective is to replace the original
function by one that serves the same purpose but is more realistic. A useful clue
is that the system is nearly always subjected to an external source of noise, which
depends on what might be called ambient noise rather than temperature. On that
basis, we can develop a model of "thermal" detrapping that remains consistent with
the spirit of the original approach. At the onset of an avalanche at angle e > e,-,
the flux of rolling species was given in the previous model by an equation of

lOWe also note, incidentally, that the present model leads to a process analogous to a first-order
phase transition, consistent with the conclusions arrived at in Section 4.2.1. This is, of course, in
contradiction with the predictions of the SOC model.
152 4. Granular Media in a State of Flow

the form

R(x, t) = Ri(x + vt, 0) exp[y(e - ee)t].

We retain the same functional form here, except that R i (x, t) now represents a
source of ambient noise due to the mechanical agitation of the system, and ee is
an angle corresponding to equilibrium. This can be justified by picturing particles
trapped in a potential well U that depends on the angle e through a relation of the
type U = mg dj(e), where j(e) is adecreasingfunctionofe which must obviously
vanish when e = Jr 12. The mechanical noise is described in terms of an effective
temperature Teff. The source of rolling species is then of the formd exp( -U 1kTeff),
which varies as d exp[a(e - em)], where em is defined as the angle at which the
detrapping potential equals the effective "thermal" energy associated with the
ambient noise. Under these conditions, we have U ~ kTeff and = mgdl kTeff,
and it can be verified that nucleation is suppressed as long as e < em. It only
begins when e = em' We should point out that in this model the angle em depends
critically on noise.
In short, the essence of this model is to give a preeminent role to the ambient
noise. This noise is responsible for detrapping the rolling species and is instru-
mental in determining the angle of movement.
De Gennes has shown that this approach accounts for both types of movement
identified in Section 4.I-a continuous flow (class A) at high velocities, and an
intermittent one (class B) at low velocities. In other words, there are always some
grains in motion in a class-A movement and, in particular, there exists a steady-
state solution for which the profile h(x, t) is constant, which is perfectly consistent
with experimental observations.
Intermittent or continuous avalanches in a rotating drum can be described by
the following system of equations

ah ah
-=-yR-+wx,
at ax
aR aR ah
-=v-+yR-,
at ax ax
where the flow is assumed to take place from left to right. Note the presence of the
wx term, describing the effect of the rotation. How to solve these equations is left
as an exercise. In case of difficulties, the reader is referred to the original article
[86]. We conclude this chapter by seeking the steady-state solution pertaining to a
class-A regime in a rotating drum of diameter 2L. The geometry is illustrated in
Figure 95.
The relevant boundary conditions are R = 0 at x = ±L. By further imposing
the restriction that the time-derivatives vanish throughout, we find steady-state
solutions of the form
W 2 2
Rss(x) = -(L - x )
2v
4.2 Avalanche Models 153

(a) (b)

-L+-_------:

-L o +L

FIGURE 95. Geometries used in conjunction with de Gennes's amended model based on
a system of coupled differential equations (see text).

and

These solutions predict a slope with a parabolic correction, illustrated in Figure 95,
although the deviation from a straight line remains small.
It is useful to remember that solutions of type A (continuous flow) and type B
(intermittent slides) crop up systematically in granular flow problems. They reflect
the metastable nature of piles whose slope evolves between the angles em and e,..
As a further example, we may want to determine the steady-state solutions when
a silo of diameter 2L is being filled from the top by a flux Q of particles falling
through a hopper. We assume that the flow is continuous and free of arch effects.
The orifice of the hopper is at x = 0, and the lateral walls of the silo are at x = ±L.
Using the same conditions as in the previous example, we find that the steady-state
solutions are

Q ( 1-
Rss(x) = 2v LIX I)
and

ess - e = (ah)
y
ax ss
-----
v
y L
I
-Ixl
Here again, the deviation of the steady-state angle from the angle of repose is
small, being only of the order of d / L.
This modified model allows a variety of situations to be analyzed, although,
at the present time, it does have limitations related to the fact that it treats a
granular as a continuous medium. One thing it cannot do, for instance, is account
for phenomena of fracturing and fragmentation (see Section 3.2.4) which, as we
now know, are likely to occur in piles subjected to any sudden change, such as
when the door of a filled silo is opened abruptly.
5
Mixing and Segregation

5.1. Introduction
Liquids have a well-known predisposition for miscibility.l Dry granular materials,
by contrast, are notorious for being difficult to mix homogeneously.2 Two granular
materials differing by their density, shape, size, or even by their micromechanical
properties (such as coefficient of elastic restitution and friction), exhibit a dis-
tinct propensity for segregation. This phenomenon is a fundamental property of
the granular state and an unending source of frustration in industry. Whenever a
mixture undergoes a flow, a vibration, or a shearing action, the components tend
to separate partially or completely, depending on the circumstances. By analogy
with chemical reactions, we may say that under the influence of various stimuli
a granular mixture inexorably tends to self-organize so as to locally reconstruct
clusters of identical particles.
As we have pointed out in the Introduction to this book, the segregation of dry
granular materials can be routinely observed even in the most primitive table-top

1Since liquids are made of particles subject to Brownian motion, thermal agitation alone produces
mixing. As pointed out in Chapter I, the Brownian motion of granular particles is entirely negligible.
Another source of energy is then required to achieve mixing. Vibration is a logical candidate, even
though the final result is often the exact opposite of what is being sought!
2The concept of "homogeneous mixture" needs to be clarified in a granular material. A mixture
composed of a fraction u: of granules A and f3 of granules B (with u: + f3 = I) is said to be homo-
geneous on a scale;" if a volume ;,,3 contains the two ingredients in the right proportions. We can
appreciate that optimum homogeneity will be achieved when the smallest scale ;", for which the
mixture can still be considered homogeneous in the sense just defined, is of the order of the size of
the largest of the two types of particle.

J. Duran, Sands, Powders, and Grains


© Springer-Verlag New York, Inc. 2000
5.1 Introduction 155

:-------.:
Wavelength A

FIGURE 96. Oyama's horizontal drum rotates around its axis. An initially homogenous
granular mixture segregates in vertical bands with a spatial periodicity A.

experiments by shaking a mixture of different grains (wheat, corn, rice, salt, are
all good candidates) in a test tube. 3 Likewise, farmers know very well that tilling
fields causes large rocks, in seemingly endless supply, to work their way up to
the surface. Peasants in India take advantage of the segregation properties of dry
granulars by shaking their harvest of chickpeas in baskets in order to separate
them from other materials. When Brazil nuts, mixed with other smaller nuts, are
transported in pick-up trucks over the rough back roads of South America, they
invariably end up on top of the load. 4 In short, the phenomenon of segregation in
dry granulars is universally recognized, even though it may not be well understood.
The first recorded observation of segregation in a three-dimensional medium
was described by Oyama in 1939.

5.1.1. Oyama's Cylindrical Drum


Oyama's experiment is schematically illustrated in Figure 96 [89]. It consists in
mixing two granulars of the same type but of different sizes and colors. For instance,
we may use glass beads, some of which are transparent and 500 /.-Lm in diameter,
while the others are dark-colored and 100 /.-Lm in diameter. They are mixed in
equal volumes and placed in an elongated cylindrical container which is rotated
horizontally around its axis.
As Oyama reported, the large and small beads divide themselves in vertical
strata, as depicted in Figure 96. The process starts with just a few slow rotations
of the drum. After a sufficiently long time, the segregation continues, first into
three, and then five zones typically observed after about an hour. The phenomenon
takes place only when the rotation speed is rather slow, although nobody really

3The importance of using dry materials cannot be overemphasized. It is essential that any interaction
with the ambient fluid be negligible. If not, the problems are of a radically different nature. Well-
designed cement trucks, for instance, are perfectly capable of adequately mixing gravel of various
sizes with cement and water.
4The example of Brazil nuts has become a paragon in matters of granular segregation following the
publication in Physical Review Letters of a paper entitled "Why do Brazil nuts are on top?". To speak
of "Brazil nut" phenomenon has become synonymous with the problem of segregation by size.
156 5. Mixing and Segregation

knows why. Spectacular as it may be, this experiment remains poorly understood. It
continues to be the object of numerous studies, some of which rely on sophisticated
techniques such as nuclear magnetic resonance (see Section 3.2.3) [90].
Be that as it may, the experiment underscores the amazing efficacy of segregation
by rotation or, to be more precise, by shearing. As mentioned earlier, segregation is
a very general property of mixtures in which particles are in motion relative to one
other. A careful observation of the mechanisms potentially involved has led some
authors to distinguish at least two different modes of granular segregation. They are:

• Segregation by Vibration. The relative motion of the particles is imparted


by shaking the container-usually in a vertical direction. When the situation
involves small particles in the midst oflarger ones, we may further distinguish
segregation by convection, by arch effects, and by percolation.
• Segregation by Shearing. Here, segregation is caused by a differential flow
of overlaid sheets, in the manner described in Section 2.4.2. A useful picture
is that of a bulldozer churning large rocks to the surface of the ground, as is
commonly observed in earth-moving projects.

Given our limited understanding of this phenomenon, it is important to acknowl-


edge that the details of the mechanisms at play remain for the most part a mystery.
No one has a handle on the fundamental laws governing the segregation of objects
with different densities or masses, or with different mechanical coefficients. That
goes both for segregation by vibration and by shearing. Likewise, very little is
known about the segregation of objects of different sizes, except perhaps in the
context of a few simulations of the type to be discussed in Chapter 6. For reasons
that are quite easy to understand, most researchers currently working on this prob-
lem have limited themselves-at least up to now-to segregation by size. That is in
fact, the primary topic covered in this chapter. As we are about to see, even this
issue turns out to be anything but trivial. In a first step, we will restrict our focus
to the segregation of a single particle differing in size from all the other particles
that constitute its environment. The purpose of this step-by-step approach is to
hopefully extract the fundamental laws governing the physics of the process. It
appears to be a general property of granular materials to want to expel the largest
objects out toward the periphery. For instance, shaking a granular medium causes
large particles to migrate toward the top of a pile.

5.1.2. Potential Energy of a Heterogeneous Pile


Before attacking more specifically the mechanism of demixing in vibrated or
sheared granulars, it is important to realize, starting from elementary considera-
tions of the energy of heterogeneous piles, that segregation by size is indeed a
phenomenon characteristic of these materials. As it turns out, this property can
be understood as a consequence of Reynolds's dilatancy principle, consistent with
arguments developed in Section 3.1.3.
We begin with a few general observations about stacks of spheres (in three
dimensions) or of cylinders (in two dimensions). Consider two spheres of masses
5.1 Introduction 157

M and m, and radii Rand r (with M > m and R > r), stacked on top of each
other. If the large sphere is on top, the potential energy E p of the stack can be
written as

E p = mgr + Mg(R + 2r),


where the horizontal base on which the edifice rests is taken as reference. Both
spheres are assumed to be made of the same material and, therefore, have the same
volumetric density. This additional piece of information leads to an expression for
the potential energy in terms of sizes only

E p ocr 4 + R 4 + 2rR 3 .
Interestingly, this expression is not symmetrical in rand R, which means that the
energy depends on whether the large or the small sphere is on top. Not surprisingly,
the configuration with the small sphere on top is energetically favorable. We may
find a degree of solace in this result. Yet, our intuition clearly fails us when it comes
to the phenomenon of segregation by size which, as already indicated, invariably
drives the largest-hence, heaviest-particles toward the top.

Superposition of Stacks-Two Compact Stacks


We next consider the stack shown in Figure 97. The structure, depicted here in
two dimensions, is made of two kinds of marble made of the same material. The
marbles differ only by their diameter and are arranged in two zones stacked in a
compact triangular lattice (which gives it maximum compactness, in accordance
with Section 3.1.3). They are placed in a cylindrical container of cross-sectional
area S. We designate by V and v, respectively, the volumes of each of the zones
occupied by marbles of radii Rand r. Assuming that the largest marbles occupy

FIGURE 97. Two types of particles A and B of different sizes are stacked in superposed
layers. In the absence of any structural defect, AlB and BIA configurations are equivalent
from an energy point of view.
158 5. Mixing and Segregation

(a) (b)

16x16 16x16
Planar square lattice Triangular lattice

FIGURE 98. Two stacking modes for two-dimensional structures. The triangular config-
uration (b), being as compact as possible, has a lower potential energy than configuration
(a). It is therefore energetically favorable and more stable.

the upper part of the container, the potential energy of the entire pile is given
by

This time, which type of marble is on top has no bearing from an energy point of
view, and there is no reason to anticipate any segregation, at least as long as inter-
face effects are absent. This result is obvious, inasmuch as two layers of identical
density are always in equilibrium.
Without the benefit of a more detailed analysis, we may begin to suspect that
structural defects in a lattice made of particles of different sizes may have some-
thing to do with the tendency of large particles to work their way back up. To
explore this possibility, we return to two configurations already discussed in Sec-
tion 3.1.3. The relevant structures are reproduced in Figure 98.
An elementary analysis of the two stacks shows that the triangular lattice is-
based on simple energy arguments-more probable and stable than its square
counterpart. As we have seen, the compact triangular lattice is the only one to
exhibit characteristics consistent with Reynolds's dilatancy principle. The reader
will recall that a compact stack subjected to any distortion can only respond by
expanding, which increases its potential energy. If we realize that distorting a
compact triangular lattice necessarily entails the creation of defects in the stack,
we may legitimately inquire whether such defects may tend to concentrate in the
lower or the upper portion of the pile. In particular, what can energy arguments
tell us about this question?

Where Are the Defects Concentrated?


We start from a structure made of identical spheres in a state of maximum com-
pactness, and assume that a number of defects are somehow created either in the
upper or in the lower region, both of which have the same volume v. Our intuition
suggests that the energetically favorable situation is to have the defects near the top.
We can verify this by calculating the potential energy in both cases. The potential
5.1 Introduction 159

Faults

Compacted
Zone

FIGURE 99. Defects are created when a large disk is inserted into a two-dimensional stack.
The photograph was obtained by back-illuminating a real stack. The lower part of the stack
remains compacted, while the triangular symmetry of the upper part is greatly disturbed by
the introduction of the large disk (after [93]).

energy is denoted E pu if the defects are in the upper region, and E pi otherwise. If
the defects involve an increase in volume dv, the calculation indicates that

E pl - E pu ex vdv > 0,

which implies that defects near the top are indeed energetically preferred. Another
way to express the same result is that, in a cylindrical container, the potential
energy is minimum when the less dense material is on top.
An equivalent two-dimensional experimental configuration, similar to those we
have discussed in Chapter 3, is easy to implement with suitable provisions for
agitation. s All we need to do is to introduce into the container a single larger disk
of the same volumetric density as the rest of the pile, and we can readily witness
the phenomenon just described. As demonstrated in Figure 99, the intruder causes
local distortions in the lattice by creating defects which tend to migrate to the upper
part of the structure. 6
When viewed in this context, the process of segregation by size emerges as one
of the consequences of the dilatancy principle. The introduction of an intruder
necessarily causes a local distortion of the lattice manifesting itself in a local
expansion. The expanded and, therefore, less dense portion of the pile tends to

5The meaning of the phrase "suitable agitation" deserves to be thought out carefully. It may be said
that shaking (or vibrating) a complex granular edifice enables us to explore perhaps not all but at least
many of the possible configurations of a pile. Several simulation approaches (notably the Monte Carlo
technique, to be discussed in Chapter 6) capitalize on this observation by minimizing the energy after
each perturbational event through various relaxation processes.
6We will often use the term "intruder" to refer to a particle whose size or other properties differ from
those of the "normal" sea of particles which, for the sake of simplicity, we will consider uniform.
160 5. Mixing and Segregation

move toward the top, dragging the intruder along with it. This would suggest that
the shape of the intruder may playa crucial role in the process of segregation,
depending on how readily it may fit in the surrounding lattice.
There are too many potential objections to accept such a crude explanation at
face value. In an attempt to get a better grasp of the phenomenon, we will have to
refine our understanding of the successive relaxation states of a pile undergoing
segregation. Until then, the one idea that should stand out and be remembered in
what follows is that size segregation implies defects created by the intruder in its
environment.

5.2. Segregation by Vibration


As pointed out on several occasions already, shaking or vibrating a mixture of
particles for the purpose of studying the process of segregation is a most efficient
and reliable method. It is routinely used both in industry and in the laboratory.
Indeed, vibrating a collection of particles is the best way to explore systematically
a large number ofpossible configurations. We might add that this mode of excitation
lends itself readily to numerical computer simulations, an advantage that is not to
be overlooked. The frequency and amplitude of the vibrations are easy to control,
and a typical experiment boils down to following what happens to a large and
clearly marked particle in its environment. By adding a few tracers to the matrix,
we can even get a fairly clear picture of how the surrounding particles move about.

5.2.1. Simulation of Segregation by Size


Before proposing a model that can account for the behavior of a vibrated two-
dimensional pile composed of a single intruder in an otherwise homogeneous
environment, it is useful to pause and consider Figure 100, as it highlights a
number of remarkable characteristics of such structures.
For starters, we note that both computer-generated and real-life stacks feature
stacking faults in the form of distortions and dislocations in the upper part of the
structure. This is fully consistent with our previous elementary energy consid-
erations. In addition, we see that the perturbation caused by an intruder whose
size is different from that of the particles composing the matrix develops in an
essentially triangular pattern. We might think that this pattern is simply a con-
sequence of the symmetry characteristic of the type of two-dimensional lattice
considered here. However, as we shall soon see, this property is not restricted to
two-dimensional lattices. Computers can serve as an inexhaustible source of syn-
thetic three-dimensional piles, whose geometry can be easily studied [91]. It turns
out that, in this case too, defects are generated in the upper part of the pile, and their
geometry is quite similar to what is observed in two dimensions. That provides con-
fidence for extending the arguments we are about to develop to three-dimensional
piles.
Lastly-and this is crucial for the model discussed next-we can see that a
large-size intruder, such as in Figures lOO(a) and (c), does not necessarily have to
5.2 Segregation by Vibration 161

(a) (b)

(e) (d)
FIGURE 100. Different configurations of a two-dimensional inhomogeneous stack. Dia-
grams (a) and (b) are computer-generated structures, while (c) and (d) are photographs of
actual real-life stacks. Note that a large intruder can be in a stable position without having
to be in contact with particles immediately below (arch effect) (after [93]).

rest on lattice lines associated with the matrix. Instead, they can be propped up
above such lines by lateral particles, somewhat like the arch of a cathedral rests on
stones that transmit its weight to side columns. Pursuing this metaphor, we refer
to this phenomenon as the arch effect or vault effect.
When trying to model the dynamic properties of such a system, we need to
inventory all possible stable positions of the intruder. Stability occurs under one
of two conditions:

• The intruder rests on a lattice line defined by the ordered arrangement of


spheres constituting the environment.
• The intruder is kept above a lattice line because it sticks at two points marked
by arrows in Figure 10 1. The line formed by joining these two points traverses
the intruder below its center of gravity. Should that process fail, the intruder
drops back down to the next lowest lattice line.

Modeling this situation involves solving a topological problem, which goes some-
thing as follows: The intruder is raised a very small step at a time, as depicted
in Figure 102, and the system is left to reorganize itself by relaxing around the
intruder. The new arrangement is examined to determine if it is stable or unsta-
ble. If it is unstable, the intruder is raised some more by a tiny amount, and the
stability is examined again. The process makes it possible to find eventually all
stable configurations.
162 5. Mixing and Segregation

FIGURE 101. Simulation of the equilibrium of an intruder by the arch effect. On the left
is a photograph of an actual stack of metal spheres containing a foreign disk (after [93]).

Two-Dimensional Model
The relevant geometry is illustrated in Figure 103(a). Let cD = R/ r be the ratio
(> 1) of the radii of the particles. As indicated earlier, the effective part of the pile
is confined within walls Bl (T) and B 2 (T), which in the present case intersect at an
angle of 60°. Our purpose is to determine all stable positions of the intruder as it is
being raised in a step-by-step fashion up the pile. The first thing we notice is that,
because of the geometry of the structure, we do not have to explore a height greater
than one period of the structure, which is given by 8 = 2r~, or 3.46 times the
radius of the dominant particles. Let h be the altitude of the intruder's center, as
shown in Figure 103(a). We keep track of each particle by its row (index i) and
column (index j). Simple geometry considerations show that the stable positions

FIGURE l02. Method for seeking the equilibrium positions in a stack by exploring different
possible configurations.
5.2 Segregation by Vibration 163

(a) (b)

FIGURE 103. Diagrams of two- and three-dimensional stacks used to develop a model of
segregation by the arch effect.

of the intruder, i.e., when it rests on a lattice line, are obtained when

where the superscript S indicates that the position is stable. The index k runs
from [(_l)i+l + 1]/2 to Int((i + 1) /2), the operatorInt(m) designating the nearest
truncated integer of the argument m.
Next we look for stable positions of the intruder via vault effects, that is to say,
when it rests on two particles located below its center of gravity. We start from
a situation where the intruder is in contact with the two lateral walls B 1(T) and
B2(T), which occurs when h vI = 2r. When the intruder is raised gradually, it will
find a new stable position when two small particles can just squeeze below its
center of gravity. This happens when h v2 = (R + 8),J3 ;:;" r,J3(cD + 2), where 8
is the space between the intruder and the lateral walls.? The fraction S of stable
vault configurations over a period e is given by

h v2 - h v1 2- ,J3
S=l- =~cD;:;"O.077cD.
e 2y3
A smooth rise through a continuum of vault configurations is obtained when S = 1,
which happens when cD~D ;:;" 12.9. As such, the quantity cD~D can be construed as
a critical ratio ofdiameters marking the boundary between two types of behavior
[92], [93]. On one side of the dividing line, the intruder rises continuously, while
on the other it goes through a series of discrete steps determined by the size of the
smaller particles. This line of reasoning makes it possible to calculate analytically,

7In this approximation, we assume 8 '" 2r. The exact solution would require lining the boundary with
small particles. The walls would then no longer be straight but made of a succession of connected
half-circles. The approximation is justified a posteriori by numerical simulations.
164 5. Mixing and Segregation

--
<I>
~ 20
'"
OJ
_. 13

-- -
.0 11
'0 15
d
E. -
I--
~ 10 1-----
--
---- -
9
7
5
o - - 3
~o t--- 1.2
c. 5
OJ
:0
.!9 (
lih )
(f) 0
o 1 2 3 4
Intruder's displacements (no. of beads)

FIGURE 104. Ascension diagram of an intruder of radius et>r in a matrix of particles of


radius r. The thick solid lines indicate the stable positions of the intruder as it rises via
the arch effect. The horizontal plateaus (fine solid lines) correspond to a discrete series of
positions in which the intruder rests on a lattice line of the matrix. The quantity oh denotes
the small relative displacements of the intruder and the surrounding lattice.

or by means of simulations which will be discussed in Chapter 6, diagrams showing


the intruder's stable positions plotted against its displacements. Figure 104 is an
example of such a plot.

Three-Dimensional Model
It is relatively straightforward to extend the preceding model to the case of three
dimensions. The triangle (T) becomes a tetrahedron which obeys the same sym-
metry. Using the notation indicated in Figure 103(b), we have
I R-r R+r
h-h = - - = - -
sin \II tan \II I '
where h = 3R, hi = 3r, \II = cos-1(1), and \II' = tan- 1eJ2/2).
The critical diameter for continuous rise via vault effects in three dimensions is
given by

",3D = 3 + .)2 i':::i 2 78


'*'c 3-.)2 .,

which tums out to be remarkably close to values found by numerical simulations,


which will be covered in Chapter 6 [91].

5.2.2. Experiments on Segregation by Vibration


The first quantitative experiments aimed at studying the phenomenon of segre-
gation by size were begun only in the late 1970s. For lack of a better technique,
the experimental approach consisted simply in measuring the time taken by an
intruder placed at the bottom of a container filled with a granular material to make
its way back up to the surface when subjected to vibrations [94]. This method was
much too primitive to reveal the subtleties of granular segregation which we are
about to discuss.
5.2 Segregation by Vibration 165

(a) (b)

2-D

1
FIGURE 105. Diagram of experimental methods used to study the segregation by shaking
of a large particle in either (a) two-dimensions or (b) three-dimensions. In (a), a large
marked intruder is immersed in a population that includes a few tracers (black particles),
whose progression is tracked by image processing techniques. In (b), the movement of
tracer particles in their environment is monitored by direct visual observation.

Modem techniques take advantage of the possibilities offered by image pro-


cessing (see Section 3.2.3), as well as nuclear magnetic resonance, which makes it
possible to monitor what goes on inside a three-dimensional opaque system. More
direct methods rely on suitably prepared samples containing tracer particles. The
technique, schematically illustrated in Figure 105, has been used to study both
two- and three-dimensional configurations [93], [95].
A typical implementation of the approach is shown in the photograph of Fig-
ure 106. The apparatus uses two small cameras. The first, placed next to a vibrating

Tracking video camera

Marked intruder

Vibration source

Camera monitoring
the vibration
amplitude

HQrlzontal trans!llltion stage

FIGURE 106. Photograph of an experimental setup designed to track the ascending move-
ment of an intruder in a two-dimensional configuration.
166 5. Mixing and Segregation

(a) (b)

_ " It. ...


.......... .. .. .. .. .. 10 ..
..
.... . ..
.. .. .. .. .. .. .. .. 4 .
..

:'I%i\\§~H~~1ii~~!~~
FIGURE 107. Experimental observations of the movement of an intruder. Diagram (a)
corresponds to a continuous rise (<I:> = 16 and [' = 1.2), while (b) reveals successive
plateaus (<I:> = 2 and [' = 1.4). The doted white line in (a) is an artifact due to the image
processing technique. It is quite real, on the other hand, in (b), which was generated by
directly displaying the height of the moving object as a function of time. The horizontal
scale corresponds to about 1 hour (after [93]).

plate, measures accurately the amplitude of the vibration. The second is mounted
on a translation stage and connected to the image processing electronics. With
thresholding and the image addition techniques described in Section 3.2.3, it is
possible to track one or more particles during the course of an experiment. 8 Alter-
natively, by moving the camera horizontally at constant speed, the vertical position
h(t) of the intruder can be displayed directly on the monitor's screen as a function
of time. We will see several examples of this technique below.

Experiments on Continuous and Intermittent Ascent


An apparatus similar to the one just described has been used to study the different
modes of ascent of cylindrical disks of various sizes immersed in a uniform granular
medium. Experiments show that a small intruder "sees" the discontinuities of
the granular environment, which is to be expected. A large disk, on the other
hand, rises smoothly without any pauses, which is not at all intuitive. Disks of
intermediate sizes exhibit continuous rises interspersed by plateaus, in agreement
with the predictions of the ascent diagram.
The model does seem to predict fairly well whether an intruder will ascend cont-
inuously (Figure 107(a)) or intermittently (Figure 107(b)). On the other hand, it
tells us nothing about what causes the ascent in the first place. A careful observation
of the vibrated container offers some hints as to what drives the intruder toward
the top.
With the aid of a suitable stroboscopic lighting system, it is easy to observe
fractures with various lifetimes appearing in the immediate vicinity of the intruder,
as shown in Figure 108.

S"Thresholding" consists in defining a particular level of brightness. Anything brighter than that level
is considered white (or I), and anything darker becomes black (or 0).
5.2 Segregation by Vibration 167

Fluctuating and reversible fractures

Long-lived
fractu res

FIGURE 108. Photograph showing the relatively random appearance of fractures around
the intruder (after [93]).

Referring to the ascent diagram of Figure 104, we can make the following obser-
vations:
• A small intruder (one that is characterized by <P < 12.9) requires a large rel-
ative displacement 8 between itself and its environment in order to overcome
each step in the ascent diagram. Accordingly, we may anticipate a sizable vi-
bration threshold to initiate this kind of motion. Experiments indeed confirm
this .
• A large intruder (one with <P> 12.9) proceeds much more readily while re-
maining continuously in an arch configuration. We can predict-and it is
experimentally verified-that the amplitude threshold is much lower than in
the previous case. In other words, a larger intruder moves up far more easily
than a small one.
As we will see, this is further confirmed by diagrams of upward speed as a function
of the ratio <P = R / r, at least on condition that we avoid the convection regime.
In this respect, experiments that make it possible to observe simultaneously the
upward drift of the intruder and the relative motions of the surrounding particles
are highly revealing.
Finally, we might point out that the phenomena of continuous and discontinu-
ous ascent just discussed have never been observed up to now in three dimensions,
even though simulations predict them. That does not necessarily mean that they
do not exist.
Convection or Arch Effect?
Experiments conducted in two or three dimensions show that, when the vibration
is sufficiently intense, particles experience a phenomenon of convection similar
to what we have already encountered in Chapter 3. We proceed to provide ad-
ditional details pertinent to these two situations. We start with three-dimensional
configurations.
168 5. Mixing and Segregation

t
lo} (b) (e)

FIGURE 109. Schematic description of three-dimensional experiments on the segregation


of a single intruder. Diagram (a) shows the initial configuration. In (b), the intruder has
begun its ascent, while some small black tracers have started to move down along the lateral
walls. Diagram (c) depicts the situation some time later, when convection recirculates the
black particles dragged to the bottom along the sides, while a central convective flow pushes
the intruder toward the top (after [95]).

Convection and Segregation in Three Dimensions


A series of experiments, conducted in a cylindrical container with and without
an intruder (see Figure 54), demonstrated convincingly the existence of convec-
tion movements in a column of granular material subjected to energetic impulses
separated by periods of the order of a second, so as to let the system relax back
between successive excitations [95]. Figure 109 illustrates schematically the kinds
of phenomena that are typically observed in such an experiment in the presence
of an intruder.
The results shown in Figure 110 demonstrate that, regardless of the size of the
intruder, its upward movement proceeds at the same speed as that of the convective
flux of the matrix. For purely geometrical reasons, large intruders cannot return
back down alongside the walls the way the main population of particles does.
We may conclude that we are dealing here with segregation by pure convection.
Segregation by convection takes place both in two- and three-dimensional config-
urations. It is very important to remember that when convection drives the process,
the ascending speed does not depend on the size of the intruder, in marked contrast
with the arch effect mechanism. 9
Convection and Segregation in Two Dimensions
We now report the results of experiments conducted in two-dimensional cells of
the type described earlier [97]. Figure III shows CPP photographs of a

9Some degree of caution is warranted in making this statement. It has not actually been proven that
the intruder itself does not somehow induce convection. The dynamic maps presented in Figure 111
would suggest that such is not the case and that the intruder does not promote convection, at least not
below its own height. This point is still being debated [96].
5.2 Segregation by Vibration 169

E 0
~
Q)
l.l
~ -5
::J
til
.8 -10
Q)
l.l
c
rn
1il -15
0 0

-200 -100 0 -200


Number of impulses

FIGURE 110. Successive positions of different intruders plotted against the number of
impulsions applied to the container. The ratio <t> is equal to 9.5 (crosses), 6 (circles), and
1 (squares), respectively. Note that large intruders remain trapped at the surface after com-
pleting their climb, while small ones (square data points) are dragged back down to the
bottom by convection (after [95]).

two-dimensional pile vibrated vertically with varying accelerations. Figure III (a)
reveals a process of convection identical to the one just discussed in three dimen-
sions. By contrast, Figure 111 (b) is indicative of an arch effect squeezing the in-
truder. As emphasized above, this phenomenon depends on the intruder's size. The
velocity of the upward movement is presumably a function of the diameter ratio <P.
This was entirely confirmed by a set of experiments recording the altitude h(t) of
intruders of various sizes as a function of time for a given acceleration and container
configuration. The results of these experiments are reproduced in Figure 112.

convection
,.... >,;'", ", ~onvection

... ...... ......;


,

:'". . .:~~~ ll:~:";:?'?:':


!~ "-"::;~ "', '. ..,:
.. arches' ,
(a) (b)
FIGURE 111. Diagram (a) shows a typical segregation mechanism by convection observed
with a relatively strong excitation (f' = 1.6). Diagram (b) was obtained with a weaker
acceleration (f' = 1.2). It reveals an arch effect mechanism that displaces markers laterally
below the intruder. This latter mode of segregation acts differently on intruders of different
sizes (after [93]).
170 5. Mixing and Segregation

I /<I>=12~9
<1>=16.3

10 .",. /<1>-10.5
," <1>=9
i) I
~
<1>=5.3
.- 8 . I
E I I
()
'-"
Q) 6 / ..
"'C
.....
::l
4
t1r.... .---lntermittent

«
;;:::; Events

2 <1>=3

0
0 10 20 30 40 50 60
Time (minutes)

FIGURE 112. Positions h(t) of intruders of various sizes immersed in a bath of particles
1.5 mm in diameter. The larger the size of the intruder, the faster its ascent (after [93]).

Here the acceleration was held constant at r = 1.25. The various curves h(t)
have been displaced along the horizontal axis to make it easier to tell them apart.
The results clearly demonstrate a process of size segregation consistent with the
arch effect model discussed earlier. Small-size intruders (<l> < 12.9) experience a
discontinuous ascending motion marked by a series of steps and plateaus. The
greater the size of the intruder, the less discontinuous its upward movement, in
agreement with the model. All of this is consistent with the ascent diagram shown
in Figure 104. Furthermore, with small enough intruders (characterized by <l> < 3),
no ascending movement occurs at all, at least not for this particular acceleration
and over the duration of the experiment (about 1 hour). The results, summarized
in Figure 113, prove that there is indeed a threshold diameter below which any

14
c
'E
......
12
E 10
.s 8
~
'u
0 6
~ 4
C
OJ
u 2
«
Ul
0
0 5 10 15 20
Diameter Ratio <l>

FIGURE 113. Ascending velocity measured from Figure 112 as a function of the diameter
ratio <P (after [93]).
5.3 Segregation by Shearing 171

ascending motion is inhibited. The existence of the threshold diameter makes sense
on the basis of the ascent diagram in Figure 104. We noted at the time that, for
weak accelerations, the fluctuations 8h of the intruder's positions can be smaller
than what is required to bridge the quantum jumps between successive steps. This
is consistent with the nonlinear behavior observed in Figure 113.
Finally, the size dependence of the ascending velocity of intruders may find use-
ful industrial applications, since it provides a means to separate particles immersed
in a granular medium. We can envision the possibility of "filtering out" components
of different sizes by a proper choice of the acceleration imparted to the cell.

5.3. Segregation by Shearing


As already mentioned, segregation by shearing, which comes about when two
sheets of granular material slide past each other at different velocities, turns out to
be surprisingly effective. The universal character of this phenomenon, involved in
situations as diverse as geophysical processes such as landslides, mixing drums,
and channeling through chutes, has prompted numerous recent investigations in
two-dimensional configurations [98], [99]. Here again, we will restrict ourselves
to the physics of the segregation of objects of different sizes, which is still in its in-
fancy, as many seemingly simple phenomena remain unexplained. Segregation by
shape, density, or micromechanical properties has not yet been studied thoroughly
enough to warrant discussion in this chapter. We will begin with the behavior of a
single intruder immersed in a uniform environment. This will be followed by an
analysis of segregation in a mixture of two granulars of different sizes.

5.3.1. A Single Particle in a Uniform Medium


As pointed out in Chapter 4, a rotating drum is a convenient tool to study the flow
properties of sheets of granular materials, and we will once again resort to this
familiar device depicted in Figure 114 [98], [99].10 This cylindrical drum with a
horizontal axis is similar to the one that served us so well in Chapter 4 for studying
avalanches. To take advantage of the image processing techniques described in
Section 3.2.3, we want to select objects with a high visual contrast. For instance,
we may use a matrix of white spheres and a single, black tracer particle, whose
movements we wish to track. We learned in Chapter 4 that when the drum is rotated
slowly about its axis, a series of more or less periodic avalanches of different sizes
is set off. The avalanches are confined to a narrow layer near the free surface of
the material, which constitutes what might be called a liquid phase. The rest of the
structure is in a compacted state that can be thought of as a solid phase. The solid

lOPigure 114 is a drawing, not a photograph taken during an actual experiment. The distinction is
significant, inasmuch as it is not obvious that the configuration depicted here can truly occur in a
real rotating drum. This particular structure was arranged so as to ensure the local equilibrium of
every disk during the stacking process. There is a fundamental difference between that and the global
stability ofthe entire edifice. An avalanche boils down to disrupting the local stability, which in turn
affects the global stability of a granular pile.
172 5. Mixing and Segregation

Rotation
Witness
particle

Roughened
inner wall

FIGURE 114. Typical experimental arrangement used to study segregation by shearing.


We track the path of a tracer particle placed in a uniform matrix.

portion of the pile remains effectively bound to the cylinder. As the tracer particle
is being dragged along by an avalanche to the bottom of the pile, it is reinserted
and buried into the solid phase. The rotation of the drum then causes it to rise back
up toward the free surface.
It is important to have a clear picture of the sequence of events. The evolution
of the intruder is illustrated in Figure 115, which shows the outcome of a real
experiment. The experiment reveals that the process of segregation toward the
center or edges of the drum occurs in the flowing region at the surface of the
pile. Based on what we know about avalanches, particularly about their statistical

FIGURE 115. Velocity diagram of a tracer particle dragged in the liquid phase and rein-
serted in the solid phase. In the present case, R < r. The intruder is seen to converge toward
the center (after [90]).
5.3 Segregation by Shearing 173

properties that we discussed in Chapter 4, we might expect that the reinsertion


points should be randomly distributed at the surface. If so, a drum should be a very
efficient mixer. As it turns out, this assumption is quite wrong, and we shall see
shortly what experiments can tell us about why that is. 11
From a practical standpoint, the question of segregation by shearing can be
couched in the following terms: Given a tracer particle of radius R immersed in a
bath of identical particles of radius r, what is the best experiment to do to find out
what causes the tracer particle to either wander over the entire space available or,
as the case may be, confine itself to a restricted portion of space, when the radii R
and r are varied?
Once again, the answer to the question is provided by image processing tech-
niques [99]. By tracking a marked intruder, it is possible to determine the rate of
occupancy throughout the half-circle filled by the granular material in the drum.
The measurement requires recording on the same image not only the various po-
sitions Pi of the marker, but also an indicator of how often each position Pi is
occupied. This is done by assigning gray levels in such a way that the more often a
site is visited, the darker the corresponding pixel appears. The technique involves
the following series of steps:
(l) An image is recorded at time tl. With suitable thresholding, the intruder is
picked out and all other particles are discarded. This produces a map with
a single black pixel.
(2) The brightness of the picture is then divided into 256, and the result is stored
in a buffer (a block of memory in electronics parlance).
(3) A new image is recorded at a later time t2. Its brightness is also divided into
256. This second image is added to the first one in the buffer. At this point,
the buffer will contain either one pixel with a brightness of 2/256 if the
intruder has not moved, or two separate pixels with a brightness of 1/256
if it has moved, in a sea of pixels with zero brightness corresponding to all
the other sites that have yet to be visited.
(4) The cycle is then repeated from the top, for a total of up to 256 times.
The lower part of Figure 116 shows the results of this type of exercise. The data is
analyzed by slicing time in discrete frames, denoted by an index i, during which
one or more avalanches may occur. If the intruder is inserted in the flowing sheet at
the radial coordinate ri at time ti, what will its coordinate ri+ 1 be in the very next
frame attime ti+ I? The corresponding correlation diagram ri+l = f (rd, referred to
as the map offirst iteration, is shown in the upper part of Figure 116. The diagram
is drawn for a region encompassing a 40° -wide sector S, bounded on the short side
by a radius R 1 corresponding to the exclusion zone of the particles in flow, and on
the long side by the inner radius R 2 of the rotating cylinder.

IIThis is an important observation. It means that segregation by shearing does not result merely from
reinsertion of an intruder during successive avalanches. A more plausible picture is for the intruder
to undergo segregation by size during an avalanche and be transported across a distance that is
determined by its relative size.
174 5. Mixing and Segregation

(a) (b) (e)

ri+l~
•• "~;.'..
ri+1J if£.'· :/;:I~':.:;~J
0.,,..,,0

~ rj ri

R/r =2/3 R/r =1 R/r =4/3


FIGURE 116. The bottom part of this figure is the superposition of 12,000 snapshots taken
every 5 seconds. It shows the spatial distribution of the sites visited by particles whose
diameter is (a) 1 mm, (b) 1.5 mm, and (c) 2 mm, placed in a bath of particles of diameter
1.5 mm. The upper part of the figure is discussed in the text. The cylinder had a diameter
of 160 mm, and its rotation speed was typically 2 degrees/s (after [99]).

The results show that, depending on the size of the intruder relative to the
majority particles, it tends to converge toward the center, to explore all the available
space, or to take refuge near the periphery. A uniformly gray area indicates near
perfect mixing.
We note first that the correlation diagrams ri+ 1 = f (ri) are all symmetrical with
respect to the principal diagonal. This observation is not insignificant. It means that
a true steady state is reached as early as the first iteration. If that were not the case,
we would see over the course of measurements a flight of the intruder either toward
large radii or toward small ones. That would translate into an accumulation of data
points either above or below the principal diagonal in the correlation diagrams. It
also means that the steady state can be described in terms of a relation of the type

IT (rj+rI rj)
IT(r;jri+1)

where P (r) is the probability of finding the intruder in region S, and IT (rj +1/ rj)
is the conditional probability of finding the particle at rj+1 when it is at rj during
the preceding frame. These results can be normalized by writing

l IT (r /r')P(r') dr
R2
per) =
Rr

and

l
R2
per) dr = 1.
Rl
5.3 Segregation by Shearing 175

5.0 - , - - - - - - - - - - - - - - - - ,

4.0
a 0.04 0'08~
0.00
-0.04
---~...... .
30
-0.08
.....- \ 0.8 1.0 1.2 1.5
\~ CD
I::" .........
Q 2.0 ''''''

1.0
~ .....c .....I:l..~:~~~~<·P-··.q· ...o/:::: ....c ...
~ '";1. ~
0.····P

, .. .. ::.::::&< " ollo ,.•••••" '"


0.0 -t-----,r--'---'::..,Ir-=---'-.--r--1----r-=-!

20.0 40.0 60.0 80.0


r (mm)
FIGURE 117. Probability ofpresence P (r) inregion S (see text). The triangles, squares, and
parallelograms correspond to particles ofdiameter 1, 1.5, and 2 mm in diameter, respectively.
They are immersed in a bath of particles whose diameter is 1.5 mm. The inset shows the
dependence of the parameter a (the inverse of the localization length) on the diameter ratio
CD (after [99]).

The experimental results have been averaged over a distance /:).r equal to three
particle diameters.
Figure 116 shows convincingly that the region of space covered by the intruder
depends on whether its diameter is larger or smaller than that of the majority
particles. The same information is conveyed by a graph displaying the probability
P (r), as shown in Figure 117. To a first approximation, the probability P (r) can
be described by a function of the type per) ex exp(ar), where a represents the
inverse of a length characteristic of the segregation process. Evidently, a changes
sign when the ratio <1> goes through unity. A positive sign indicates that the intruder
flees toward the periphery, a negative sign that it tends to converge toward the center.
Without pursuing the analysis of experimental results any further, we may note
that our avalanche model discussed in Chapter 4 is not completely accurate. In
particular, we had claimed that an avalanche, being of random size, should reinsert
a particle anywhere along a flowing sheet. Actually, as revealed by the maps of
first iteration or by tracking an individual marked particle of congruent size, both
the center and the periphery of the cylinder are attractors for the dynamics of
the system. If a particle is introduced near the periphery of the cylinder, it will
tend to stay there. Likewise, a particle introduced near the center will tend not to
wander off vary far. This implies a degree of correlation between trapping events
taking place within avalanches, somewhat consistent with the model developed in
Section 4.2.2. We might conclude that the dynamics of size segregation is governed
by two attractors, one at the center, and the other on the periphery of the cylinder.
According to this hypothesis, segregation in a rotating cylinder could be construed
176 5. Mixing and Segregation

FIGURE 118. Schematic diagram ofthe segregation process of a two-dimensional granular


mixture. The black particles are smaller than the white ones. They begin to collect near the
center of the drum after just one revolution.

as a mechanism favoring one attractor over the other. Such situations often involve
the phenomenon of bistability. 12

5.3.2. Segregation of Two Populations ofParticles


of Different Size
We now consider a mixture of two distinct populations of particles. The first ques-
tion that comes to mind is whether the observations just made in the case of a single
intruder in a sea of identical particles can be extrapolated to predict the behavior
of our mixture of two materials. In other words, can the process of separation in a
binary mixture be treated as a succession of independent steps affecting individual
particles, such as we were dealing with in the preceding section? The answer is
not at all obvious, as we are about to find out. We will show that the kinetics and
geometry of the segregation process depends on the shape of an incipient cluster
growing with a fractal structure.
We now define the problem more rigorously. Let N A be the number of particles
of diameter d A, which is to be mixed with, or segregated from, a number N B
of particles of diameter dB. As usual, we define a ratio <P = dA/d B. The entire
system is placed in a two-dimensional rotating cylinder of the type used earlier,
which amounts to a simplified two-dimensional version of Oyama's drum. The
experiment starts by mixing particles A as completely and randomly as possible
with particles B. These may be, for instance, black disks mixed in with slightly
larger white disks, as depicted in Figure 118.
The experiment reveals that after just a few turns of the cylinder, the smaller
particles have gathered in the central part of the drum. The connected mass cre-
ated this way is called the reference mass. 13 Its surface area, reached in principle

12 Abistable system has two equilibrium states. It can switch from one to the other under the effect of
some external perturbation.
13 A mass is said to be "connected" when its particles actually touch each other.
5.3 Segregation by Shearing 177

after an infinite time-in reality, this time is quite short, as we will see-is de-
noted Soo.
We proceed to highlight a few general ideas relevant to this situation. A more
detailed analysis can be found in [90]. In order to characterize the state of the
mixture, it is useful to introduce a parameter describing its degree of order. Let
S Ct) be the surface area of type- A disks absorbed in the reference mass at time t. It
is clear that S Ct) has to be smaller than Soo. The degree of segregation is quantified
by a parameter aCt) defined as
Set)
aCt) ==-.
Soo
At this point, it is natural to introduce an ordering parameter poet) that can vary
between 0 (for a completely random and homogenous mixture) and 1 (for a fully
developed reference mass). This parameter is defined in terms of aCt) by

P (t) _ aCt) - a(O)


a - 1 - a(O) .

Po (t) can easily be determined by means of imaging processing techniques of the


type described in Section 3.2.3. The results presented in Figure 119 shed some
light on the kinetics of the process of segregation.
The graph reveals two important findings:
• The growth of the ordering parameter is surprisingly fast. With a rotation
velocity as low as 1.3 rpm, the reference mass is virtually fully developed
after only about 100 s, which corresponds roughly to two full turns. The
time constant T in Figure 119 is of the order of 0.7 revolution. These results
underscore the remarkable efficiency of segregation by shearing.

1.0

-
~

C....L 0.8

"*E~ 0.6

O;~o
I1l o
a..
....Q) 0.4 o 0 00
rn
~ 0

"E 0

o .Q
Time constant 'r
0.2
Time
0.0
o 200 400 600
Time t (seconds)

FIGURE 119. Order parameter PaCt) plotted against time for a mixture of disks with
diameters of 6 and 10 mm. Thirty percent of the surface is occupied by the smaller disks.
The inset shows the dependence of [1 - PaCt)] on time Cafter ref. [90]).
178 5. Mixing and Segregation

• The ordering parameter grows exponentially. The inset in Figure 119 shows
that it evolves as Po(t) ex 1 - exp( -t Ir), which describes a kinetic process
of first order. This remarkably simple behavior remains totally unexplained.
When studying mixtures of different concentrations of particles A and B, the
conclusion is that the time constant T is practically independent of the composition.
As of yet, this too is unexplained.

Segregation Speed and Particle Size


The ratio of the diameters of the two types of particles should play an important
role in how fast the central mass forms. Indeed, we know the result in two extreme
cases:
• When the ratio <D is quite small (typically less than 0.2), which happens when
one type of particle is very much smaller than the other, a phenomenon known
as sifting takes place. It describes the ability of the smaller particles to snake
their way through the interstices between the larger ones. This situation can
be modeled using the concepts of directed percolation. The small particles
then remain trapped at the interface between the liquid and solid phases which
defines the flowing sheet. Since the medium in which the small particles evolve
is defined by the gaps between the large ones and is, therefore, independent of
their own size, the ratio <D should have no effect on the velocity of segregation.
Under these circumstances, the time constant r must be independent of <D.
• If the ratio <D is such that <D = 1, which amounts to dealing with a single
type of particle, the arguments developed in Section 5.3.1 apply. In particular,
we established that all particles roam through the entire space available. By
marking some of the particles, it was subsequently established that particles
actually tend to remain in the vicinity of the attractor nearest the point of
insertion. If so, it is, of course, impossible to cling to the definition of the
parameter T. It must diverge to infinity since, from a macroscopic point of
view, Po is required to remain constant in this case.
Numerous experiments have shown that between these two extremes, the time
constant r varies approximately linearly with the ratio <D. If r is expressed in
numbers of rotations, then we have the simple linear relation

T ~ 1.2<D with <D E [0.2,0.8].

This expression does indeed give r ~ 0.7 when <D = 0.6.


We conclude this section, devoted to segregation by shearing, with a discussion
of the surprising-and so far unexplained-dependence of the time constant on
the rotation velocity of the drum which, up to this point, has been held constant at
1.3 rpm.

Segregation Speed and Rotation Velocity


First off, we note that a rotation speed of 1.3 rpm, chosen for the experiments
described above, causes a continuous flow along the tilted surface in the drum.
5.3 Segregation by Shearing 179

This is entirely consistent with the results reported in Section 4.1, namely, that
the flow switches from discontinuous to continuous at around 0.3 rpm. Second,
the speed of 1.3 rpm is well below the 12 rpm needed for the onset of inertial
effects.
Both the previous experiments and the present one were done with 600 small
disks 6 mm in diameter, and 720 large disks 10 mm in diameter, giving area ratios
of 25% and 75%, respectively. Two surprising results emerged. First, the time
constant remained unchanged at 25 s as the rotation velocity varied from 1.3 to
8 rpm. Second, at velocities higher than 8 rpm, segregation disappeared altogether
and the mixture remained substantially homogeneous, even over periods of several
hours.
These results are completely unexpected and, for the time being, without ex-
planation. We would normally anticipate that the process of segregation should
become more efficient as the small disks pass more frequently through the portion
of granular material flowing down the slope. In fact, nothing of the sort happens.
Quite on the contrary, segregation proceeds at the same pace even though the
number of crossings through the liquid phase varies by more than a factor of 6.
Furthermore, since segregation was shown to behave as a kinetic process of first
order, we would expect it to depend monotonically on rotation speed, rather than
to suddenly drop to zero at 8 rpm.
Perhaps a plausible interpretation can be proposed if we go back to the principles
we invoked to explain the role of arch effects in the phenomenon of segregation
by vibration. We argued that a granular system must have enough time to relax
between excitations in order to adapt to the intruder's geometry. Only then can the
intruder migrate efficiently. In other words, segregation is sensitive to the size and
geometry of the objects involved and requires a finite amount of time to manifest
itself. When viewed in that light, the results described above may not be so puzzling
after all. As the rotation velocity increases, the granular surface flow becomes too
fast and chaotic, leaving too little time to adjust to the geometry of the particles.
We may even push the argument a step fmther and envision that the flowing
sheet, which is liquid-like at low speed, gradually turns into a gas as the number
of collisions between particles increases with speed. This hypothetical "phase
change" between liquid and gas may occur rather abruptly and could very well
explain why segregation suddenly disappears above a certain rotation velocity (in
the present case, at 8 rpm). Between 1.3 and 8 rpm-the regime where segregation
proceeds efficiently-the experimental results show that T increases linearly with
rotation velocity (provided that T be expressed in number of revolutions).

Fractal Growth of the Central Cluster


As we noted earlier, the central cluster made of the smaller particles is practically
formed in a very short time corresponding to two revolutions of the cylinder, or just
three times the time constant T. Having determined that, for all practical purposes,
the formation of this cluster is governed by a first-order law, as are many other
growth processes, we set out to try to understand the phenomenon on the basis of
geometrical considerations.
180 5. Mixing and Segregation

(a)

Boundary line

(b)

FIGURE 120. Diagram (a) shows particle clustering after the drum, filled as indicated in the
text, has rotated for 300 s. Diagram (b) shows a portion of the boundary of the accumulated
cluster.

There exists a prolific literature on the subject of materials synthesis by a va-


riety of growth mechanisms, including thin-layer deposition, Diffusion-Limited-
Aggregation (or DLA, for short), [lOO]-directed percolation, and several others.
Quite often, these various growth mechanisms lead to self-similar geometrical
forms, that is to say, structures with fractal properties.
Defining the segregated cluster as that delineated by the greatest possible number
of connected points, we designate by M (r) the length of the jagged line that forms
its boundary. The objective is to calculate this length as a function of the radius r
of a circle centered on a particular point of the line, as shown in Figure 120. Let
M (rs ) be the length obtained when the radius r of the circle matches the radius r s
of the small disks composing the segregated mass. We can then generate diagrams
of the quantity log[M (r) / M (rs )] plotted against log(r/ r s )' The result is shown
in Figure 121. The graph reveals that the boundary of the segregated cluster has
a fractal structure with a dimension d = 1.62 ± 0.2, which allows us to write the
functional dependence M (r) ex rd.
As it happens, numerical simulations have also been done, although in a some-
what different context [l 00]. Based on a model of directed and uncorrelated growth
of a two-dimensional structure, they gave an exponent of 1.76, not too different
from the result reported above. It can be shown, incidentally, that the exponent
should decrease when finite-size effects are present, which is most certainly the
case for our segregated cluster. To be sure, these considerations are only semi-
quantitative. Although they have great pedagogical value, the results described
5.4 Segregation in Oyama's Three-Dimensional Drum 181

a
0.0 0.4 0.8 1.2 1.6
log(r)

FIGURE 121. Normalized length M plotted against the radius r of the measurement circle.

here have yet to be confirmed by more careful experimental work and further
theoretical analysis. 14

5.4. Segregation in Oyama's Three-Dimensional Drum


As noted at the beginning of this chapter, the experiment reported in 1939 by
Oyama involves a phenomenon that remains without explanation [89]. A schematic
diagram of the experiment is reproduced for convenience in Figure 122. The reader
will recall that an initially homogeneous mixture of two kinds of particles with
different sizes gets segregated in vertically separate regions as the drum rotates.
The physics of this phenomenon has so far defied analysis.
We will first describe a few experimental observations reported in the literature
[90], [101]. We will then sketch the broad outlines of a model recently proposed
by S. Savage [102].

5.4.1. Experimental Observations


The experiments described here are done by filling one-third of the volume of a
glass cylinder 70 cm long and 10 cm in diameter, whose internal wall is lined
with roughened spheres. The mixture introduced into the cylinder is composed
of 50% of colored spheres 1 mm in diameter, and 50% of spheres of a different
color and 3 mm in diameter. The cylinder is then rotated around its horizontal

14De Gennes has recently developed a model for segregation via avalanches [43]. The model is based
on a set of coupled variables discussed in Section 4.2.3. Although it deals with a situation that is
different from a rotating drum, it leads to power laws with fractional exponents which may well have
a bearing on the experiments described here.
182 5. Mixing and Segregation

~
Wavelength A

FIGURE 122. Three-dimensional segregation in Oyama's drum. An initially homogenous


mixture of two different types of particles undergoes segregation in a number of bands
perpendicular to the axis of rotation.

axis at speeds ranging from 15 to 65 rpm. After 10 or 20 min, the mixture has
separated in bands characterized by a wavelength A, as depicted in Figure 122.
The most salient features of the experimental results can be summarized thus:
Between 15 and 65 rpm, there is little or no dependence of the wavelength on the
angular velocity. When the rotation speed drops below 15 rpm, segregation ceases
altogether, and the smallest bands tend to be the first ones to vanish.
Several authors have reported that the steady state, reached after long periods,
consists of three bands.

5.4.2. Savage's Model


S. Savage has proposed a phenomenological model based on an observation we
skimmed over briefly in Section 4.1. It refers to a phenomenon that is observed
fairly frequently but is poorly understood [103], [104]. Specifically, when the drum
is rotated at a speed sufficient to generate a continuous flow, the tilt angle of the
surface sheet with respect to the horizontal depends on the size of the particles.
Whether this reflects a kinetic drag phenomenon or, as is more likely, a finite-
size effect typical in this type of experiment, the fact the kinetic angle changes
according to the type of particles will induce a lateral flux (i.e., parallel to the x-axis
in Figure 122), which will depend on the ratio of the particle diameters as well
as on the rotation speed. IS With this observation in mind, we consider a mixture
of two populations of spheres A and B. Let eA and eB be their respective kinetic
angle at a given rotation speed. We denote by CA (x) the local concentration of
species A at a point of abscissa x. We expect the kinetic angle e(x) of a mixture of
particles to be a weighted average of the kinetic angles of each individual species.

15 Curiously,
the dependence of the kinetic angle on particle size has apparently never been observed
in two-dimensional geometries, perhaps because the total number of particles is then too limited.
Expanding on an idea advanced in Section 4.1 in connection with avalanches of varions sizes, we
may surmise that the aspect ratio manifests itself only when the number of particles is large enough,
regardless of dimensionality.
5.4 Segregation in Oyama's Three-Dimensional Drum 183

This we write

with

Focusing our attention on type-B particles, their flux will result from two com-
peting effects. On the one hand, there is a flux <PBx (,6,,8) due to the difference
in kinetic angles, which tends to drive particles along the direction of the x-axis
(see Figure 122). On the other, there is an opposing flux <P BD due to diffusion and
described by Fick's law involving a diffusion coefficient D. The total flux <P Bx of
B-type particles along the x-axis is then given by
aCB aCB
= -,6,,8-- - D--.
<PBx
ax ax
To summarize, the horizontal flux is created by differences in the kinetic angle, but
it is opposed by a diffusive component that tends to equalize the concentrations.
The formation of bands would result from competition between the two effects.
6
Numerical Simulations

6.1. Introduction
Numerical simulations aimed at modeling various aspects of the physics of gran-
ular materials, which we have touched upon throughout the earlier chapters, have
a twofold objective. l On the one hand, there are pressing incentives to solve a
number of practical problems related to the treatment of granular matter in indus-
try. Whether the issue is pesky segregation, blockages of flows by arch effects,
or disruptive internal convection phenomena (see Chapter 1), the requirements
of the industrial sector are many and, needless to say, almost always immediate.
The urgency of industry's needs and the increasingly rapid developments of cre-
ative numerical simulation techniques have prompted many researchers to devote
a great deal of effort to devising algorithms suitable for describing the behavior of
granular materials. 2
On the other hand, numerical simulations are of considerable interest from a
more fundamental point view as well. They offer the possibility to explore the
effect of many parameters which are simply not accessible to experimentation. 3
In that sense, numerical simulation has truly become an integral part of basic

1 An excellent introduction to the topic of numerical simulations of granular materials can be found in
[59].
2 At
the present time, the number of researchers engaged in numerical simulations in this particular
domain of physics substantially exceeds those pursuing experimental work.
3That is the case, notably, for the coefficients of elastic restitution E and the coefficients of friction
fL, which can be valied at will on a computer, whereas nature offers the experimenter a very limited
range of choices.

J. Duran, Sands, Powders, and Grains


© Springer-Verlag New York, Inc. 2000
6.1 Introduction 185

research. Indeed, it does not operate in a vacuum. Comparing experimental results


and numerical simulations constitutes the ultimate test to validate-or reject, as
the case may be-tentative models.
The goal of computer modeling is as ambitious as it is clear. Starting from the
properties of the elementary granules that make up the material of interest, and
incorporating whatever may be known about their mutual interactions, the objective
is to devise computational methods flexible and general enough to predict the actual
behavior of a real granular system in a variety of situations.

6.1.1. The Challenges ofNumerical Simulation


The comments we have just made could apply just as well to any other area of
the physical sciences. Yet, the case of granular materials is rather unique. We have
already hinted in Chapter 2 at the difficulties physicists face when trying to model
collisions and friction between solids. Modern numerical techniques can deal with
so-called n-body problems, even when n is quite large. That is no longer an obstacle.
The main challenge is to incorporate in the formulation of these problems the basic
micromechanical properties describing the relevant interactions, and to do so as
accurately as possible. Many decisions have to be made while setting up a model.
What is the duration tc of a collision between two particles? What is the penetration
distance (Section 2.2.2)? How should the proper time step be chosen in relation to
both the time interval between two successive events and the collision duration tc
(Sections 2.2.2 and 3.2.1)? What is the best way to model Coulomb's laws of dry
friction (see Section 6.4)? Perhaps the toughest problem of all is how to account
correctly for the effect of microcontacts and their erosion over time, as mentioned
in Section 2.2.1. Is there away to include phenomena of wear and tear and strain
hardening in the model? All these questions have to be answered before devising
a realistic simulation.

6.1.2. The Different Simulation Methods


Given this lengthy list of problems inherent to granular materials, it is hardly
surprising that a great many different strategies have been proposed over the last
few years. As it turns out, each simulation method has its advantages and draw-
backs. Whether it be in terms of computational time or accuracy of results, no
approach developed to date can satisfy all requirements. Compromise is the rule,
as each method suffers from some degree of limitation. Under these circumstances,
choosing one over another requires a good deal of caution.
Coming up with a logical classification scheme of such diverse simulation tech-
niques is no easy task. It would be a bit presumptuous to attempt a comprehensive
review of everything we presently know on the issue of simulating granular matter.
Without exceeding the scope of this book, though, it is possible to extract from
simple observations a few general principles which we will build on in the remain-
der of this chapter. Tllis approach will enable us to clarify a number of notions
which those doing simulation work routinely rely on.
186 6. Numerical Simulations

Hard Spheres and Soft Spheres


The first idea that comes to mind to simplify the level of difficulty in modeling
solids in collision is to treat them as hard spheres. In the context of numerical
simulation, however, the word "hard" does not necessarily imply that the collisions
are perfectly elastic. It simply means that there is no interpenetration or deformation
during impact, which is considered infinitely brief. The loss of linear momentum
is characterized solely by means of the coefficient of elastic restitution, at least
when rotations are neglected. That was in fact the approach we used to model a
column of spheres in Section 3.2.1.
The hard-sphere approximation is at the basis of the so-called "collisional" or
"event-driven" (ED) models, as we will see later on. It is also the principle behind
various pile-synthezing methods, including dynamics ofcontacts (see Section 6.4),
Monte Carlo, and steepest descent.
In this approach, the mechanisms of restitution of elastic energy and friction are
treated as if they were completely decoupled. Dry friction is generally modeled in
terms of Coulomb's laws as presented in Section 2.2.1.
The soft-sphere approximation is based on an entirely different principle. Here,
friction and elastic restitution come into play only when spheres penetrate into each
other, and the magnitude of the interaction depends on the penetration depth. The
prototypical algorithm in this category is the molecular dynamics (MD) model.
The essence of this approach revolves around the deformation of spheres. As such,
how long they remain in contact is of paramount importance.

Duration of Collisions and Chronology Problems


Predictably, the order in which the calculations are performed must be selected
with care. Two possibilities exist, each with its own advantages and disadvantages:

• The first option consists in sampling the system at regular intervals, using
a time step small enough to avoid "missing" an event, which could com-
pletely change the subsequent chain of events. However, as we have already
pointed out, the duration of contact between hard spheres is infinitesimally
short. Several collisions may occur in rapid succession and possibly set up
oscillations at a rate exceeding the sampling frequency, as suggested in Sec-
tion 3.2.1 and illustrated in Figure 123. Therefore, a sequential algorithm is
not a wise choice in the hard-sphere approximation. It would be more suitable
for a molecular dynamics model, where collisions have a finite duration.
• Researchers have devised an algorithm in which the timing of the sampling is
governed not by a fixed external clock, but by events themselves. This type of
event-driven approach guarantees that no event will be missed. On the down-
side, there is a risk of getting trapped in situations where a particular event,
such as an oscillation triggered by a collision, lasts for a very long time. It
then becomes necessary to rely on some test criterion to get around the trap.
We have already mentioned such a criterion in Section 3.2.1, known as LRV
(Largest Relative Velocity).
6.1 Introduction 187

Loss of
Sequential Event
Method II I I I I I I
?
Events 11111
Time

Jammed
Event-Driven Point
Method 1111

Events 11111
Time

FIGURE 123. Schematic illustration of the type of difficulties encountered in numer-


ical simulations using either sequential algorithms (top) or event-driven algorithms
(bottom).

Summarizing, ED-type algorithms are particularly well suited to describe hard


spheres, while MD algorithms, based as they are on soft spheres, are more com-
patible with regularly spaced periodic sampling. The remainder of this chapter will
provide more details on these two techniques, which have become widely used.
Next, we will describe briefly some approaches based on the mechanical properties
of contacts between solids, coupled with specific convergence criteria. This latter
approach has recently scored some remarkable successes in modeling the dynam-
ics of granular structures in a variety of situations. In connection with the Brazil nut
problem, we will describe the Monte Carlo technique, as well as the method known
as steepest descent. Finally, cellular automaton models have also been proposed
recently to simulate the dynamics of granular materials. They are close cousins of
those used so successfully in solving certain hydrodynamics problems. This area
is currently in rapid development, and it would be premature to cover it in great
depth.
Before embarking on an analysis of these various techniques, this is an ap-
propriate opportunity to provide at least a partial answer to a question we have
raised on several occasions in the preceding chapters: How to make the transition
from a discrete representation-which a numerical simulation is by essence-to a
"thermodynamical" description of a granular medium?

6.1.3. The Transition from a Discrete to a Continuous Description


We have emphasized in Chapter 3 the difficulties faced by continuum descrip-
tions when applied to a real granular medium, which is inherently discontinuous.
Predictably, differential equations become increasingly inadequate as the number
of particles involved diminishes. This type of problem was particularly evident in
our analysis of granular avalanches in Chapter 4.
188 6. Numerical Simulations

Numerical simulations modeling a sequence of events such as collisions and


friction determine at any given instant the position Xi and velocity Ui of every
particle. On the other hand, we know that classical thermodynamic theory rests on
continuous and differentiable variables such as density p, collective drift velocity v,
and macroscopic temperature T. The question is whether there exists a connection
between the two types of description. Given a complete knowledge of the positions
and velocities of all the particles (provided by a numerical simulation), is there a
way to define the thermodynamic quantities p, v, and T?4 Symbolically, we may
write

There are clearly several possible answers to this question, some more realistic
than others. The technique we describe next has the merit of being fairly intuitive.
It is rooted in the notion of "cloud," which effectively spreads out the mass of each
particle over a region larger than its actual volume [57]. In this picture, the clouds
associated with two particles can overlap, which ensures a continuous passage
from one to the other. The cloud function her) must satisfy several conditions.
They are

1 00

h(r)2:rrr dr = 1, (6-1)

her) ---+ 0 when r ---+ 00, (6-2)


her) :::: O. (6-3)

Equation (6-1) is the normalization condition, written here in two dimensions,


(6-2) states that a cloud is primarily localized around its associated particle, and
the inequality (6-3) ensures that the density p and temperature T will be positive
quantities. To simplify the calculations, we adopt a Gaussian distribution for the
cloud function h (r)

(r
her) = -1-2 exp - - ) ,
2:rr 0' 20'2
2

where 0' is larger than the diameter d of a particle (for instance, 0' = 6d) and deter-
mines the extent of the clould. This enables us to define a density p, a macroscopic

4We have already defined in Section 4.2.1 a granular temperature as the driving force of a thermal
agitation by vibration which causes the detrapping necessary to set off an avalanche. It is not in the
least proven that this definition is identical to the one we are about to put forth in this section.
6.2 Simulations of Collisions 189

velocity v, and a temperature T, through a series of three successive equations


N
p(x) =m I>(lxi - xl),
;=1
N
p(X)V(X) = m I>;h(lx; - xl),
;=1

p(x)T(x) =m L N

;=1
U2
--'-h(lxi -
2
xl) - p(X)--,
V 2 (X)

2
where N is the number of particles used in the simulation, and m is the mass of
each. These macroscopic quantities are continuous in space and time. It is pos-
sible to calculate their gradients, and as such they can be viewed as the usual
thermodynamic variables.

6.2. Simulations of Collisions


6.2.1. Introduction
As we saw in Chapter 3, event-driven (ED) methods consist in establishing a set of
general equations describing the dynamics of the system (for instance, Newton's
equation, as spelled out in Section 2.2.2). Knowing the variables {x;, Vi} for all
particles during a particular event, the next chronological event can be predicted,
and the procedure is repeated sequentially. Section 3.2.1 described the use of
this technique to model a one-dimensional pile of hard spheres in vibration. We
subsequently saw the same technique applied to problems of decompaction and
self-organization (Section 3.2.4). We will not dwell on this method any further.
The interested reader is encouraged to go back to the cited sections to review the
details. What we will do here is describe a procedure to get around problems of
accumulation which were mentioned before.

6.2.2. One-Dimensional LRV Procedure


The so-called LRV procedure (for Largest Relative Velocity) is useful when gran-
ules in a multiparticulate system come in contact and form what we have referred
to as blocks [59]. The power of the technique lies in its ability to avoid infinite com-
putationalloops that arise when the particles, assumed to be made of hard spheres,
remain clustered. The algorithm relies on a logical test to make predictions about
the outcome. As such, it avoids situations that would result in impractically lengthy
calculations. 5 Figure 124 illustrates a specific case.

5 Researchers doing numerical simulations frequently resort to such predictive techniques.which consist
in bypassing accumulation points leading to endless loops by deciding ahead of time the state a system
will find itself in. This type of trick, as it were, speeds up the computation time. It only works, of
course, to the extent that the predictions are indeed correct.
190 6. Numerical Simulations

FIGURE 124. Evolution of a group of five spheres, initially grouped in two blocks, for
E = 0.8 (after [59]).

Consider five spheres originally arranged in two blocks of two and three mem-
bers, respectively, about to collide. What is the trajectory of the five spheres after
the collision? Do the spheres coalesce into a single block? If not, do they separate
and perhaps rearrange themselves in a different pattern? Before solving this prob-
lem, it is useful to recall the definition given earlier of a "block" in the context of a
numerical simulation. Two colliding particles form a block if their relative velocity
is smaller than a predetermined value V c chosen according to the characteristics of
the computer used. From a practical standpoint, the velocity differences between
all pairs of adjacent particles forming a block have to be computed at every instant
identified by the ED algorithm. Let 1'..vi = Vi-l - Vi designate those differences
for all adjacent pairs of a block. Pairs for which 1'.. Vi < 0 do not collide, while those
for which 1'.. Vi > 0 are likely to. The LRV procedure works as follows:
(1) We pick the adjacent pair with the largest value of 1'.. Vi , or 1'..Vj = max(1'..vi),
at the moment of impact, and we let the particles (j, j - 1) collide.
(2) Collision matrices of the type described in Section 3.2.1 are used to calculate
an updated set of differences 1'.. Vi.
(3) The previous two steps are repeated until all the differences 1'.. Vi become
either smaller than V c (in which case the corresponding particles form a
block) or negative (in which case they fly apart).
It has been shown that this type of predictive approach does indeed lead to the
same result as the conventional ED method.

6.3. Molecular Dynamics (MD) Simulations


The so-called molecular dynamics (MD) methods benefit from a large menu of
algorithms developed and tried in many different test cases. It is fair to say that
this technique has matured to the level of an indispensable tool to simulate many
aspects of the dynamics of granular materials, provided that certain precautions
concerning time and spatial scales be taken. The technique relies on the concept
6.3 Molecular Dynamics (MD) Simulations 191

of soft spheres and on a sequential calculation. The primary difference with event-
driven methods is that in the present case the duration tc of a collision is not zero.
The principle behind MD methods is to solve in regular incremental steps the
equations governing the changes in linear and angular momenta of the colliding
particles. The objective is to solve the following vector equations

!1p = !1(mv) = mv - mvo = 10t' Fcrndt (6-4a)

and

!1(Iw) = fw - fwo 10t' (r x F)dt, (6-4b)

where f is the moment of inertia of the solid around its axis of rotation, r x F is
the torque exerted by the force F, and Fern is the component of the force acting on
the center of mass. Once again, we emphasize that this strategy is quite different
from the one followed in ED models, which starts from the equations governing
momentum exchanges, in the manner described by (2-2) and (2-3). In the present
case, solving (6-4) requires a knowledge of the forces F and Fern, of how they vary
in time, and of the duration tc of the collision. As a prerequisite to any molecular
dynamics simulation, it is essential to model as exactly as possible the forces
of elastic restitution and friction involved during collisions between particles. We
have already stressed on numerous occasions how fundamentally difficult this task
can be (see, in particular, Section 3.1.1), due to the inherently indeterminate nature
of the equilibrium forces in a granular stack, as they depend on its prior history, or
to our limited understanding of contact interactions between solids. This explains
the many forms of equations proposed by various researchers working on this
problem. The situation is not unlike that discussed in Section 4.2.2, in which we
reviewed a variety of functional dependences of the friction forces on velocity.
The goal of the next few paragraphs is to discuss the different types of behavior
we might encounter in modeling the contact forces F and Fern, which feed directly
into (6-4).

6.3.1. Elastic and Friction Forces


Linear and Nonlinear Equations
Consider a set of N spherical particles of diameter d i , where the index i runs from
1 to N [59]. If the particles are all identical, we obviously have d i = d for all
values of i. We can also envision without any difficulty a distribution (for instance,
Gaussian) of diameters di centered on a value d and with a spread (!1d). Let rij
be the distance between the centers of two particles of index i and j. In accordance
with Signorini's conditions, which are widely used in matters pertaining to the
mechanical properties of contacts (see Section 6.4), the forces of contact come
into play only when d; + d j < 2rij. When this condition is verified, three different
192 6. Numerical Simulations

contact forces are involved, at least under the assumption that angular momenta
can be neglected. In simplified vector notation, these forces are:

• A force of elastic restitution, related to the elastic energy stored during the
penetration of the two particles. This force is given by

(6-5)

where llij is the unit vector along the line connecting the centers of the parti-
cles i and j. This is simply the usual relation for the deformation of a spring
with stiffness K. It is obviously linear and, as such, incompatible with Hertz's
penetration model (Section 2.2.2), which predicts a power law with an expo-
nent of ~ to describe how the force depends on the penetration distance. To
allow for this nonlinearity, (6-5) is modified to a slightly more general form

(6-6)

where f3 = ~ in the Hertz model, and f3 = - ~ in the case of a soft crust (see
Section 2.2.2).
• A friction force which opposes the rupture of contacts. It plays a dissipative
role similar to that of the Euler-Coulomb dynamic friction. For generality,
two components are distinguished. The normal component is

(6-7)

where mij is the reduced mass of the system of two colliding particles i
and j, V;j is the difference between their velocities, and D n is a dissipation
coefficient characterizing the separation of contact along the direction of llij.
Likewise, the tangential component of the friction force is

(6-8)

where tij is a vector tangent to the contact, that is to say, perpendicular to llij,
along the slip direction, and D t describes the corresponding dissipation.

Here again the linear approximation contained in (6-7) and (6-8) is sometimes too
limiting. The equations are often generalized in the form

valid only when ~ (d; + d j ) > rij.


It is extremely important to understand that dissipative processes introduced
in these equations to model friction are inherently dynamic in nature. Indeed,
these equations do not account for Coulomb's static friction forces. The present
model applies exclusively to a dynamic analysis of granular piles, as does the ED
technique discussed earlier.
6.3 Molecular Dynamics (MD) Simulations 193

D
Oashpot

FIGURE 125. Mechanical model simulating contact interactions by a coupled spring-


dashpot arrangement. The dashpot acts as a shock absorber.

Mechanical Analogies
The preceding equations were introduced purely phenomenologically. They may,
however, be interpreted in terms of more concrete models that give a physical
meaning to the parameters figuring in the equations. The simplest analogy is de-
picted in Figure 125. It features a spring (simulating elastic restitution) coupled to
a linear dampener. 6
Such a simple system obviously cannot account for the subtleties of contact
interactions, such as the plastic deformations that typically occur when two collid-
ing spheres penetrate each other, as pointed out in Section 2.2.2. More elaborate
variants have been proposed to simulate these more complicated effects [38]. An
example is depicted in Figure 126. With enough creative imagination, other ar-
rangements can undoubtedly be contrived, but we should keep in mind that such
mechanical analogies have limitations and remain crude pictures of reality. Figure
126(b) describes the behavior of the system shown in Figure 126(a). The spring of
stiffness K 1 is compressed, simulating the two particles colliding and penetrating
each other. On Figure 126(b), the operating point moves up along the straight line
of slope K 1 until its abscissa is equal to et. At that point, the ratchet mechanism
jumps down one notch, which causes the stiffness to suddenly increase to K2.
If the system is allowed to relax in its new configuration, the operating point
moves back down along a different line of slope K 2 , until it reaches the point of
abscissa eto. Since the force has now returned to zero, the system is clearly left in a
different state relative to what it was, which is consistent with the phenomenon of

6 Such a dampener is sometimes referred to as LSD, for linear spring dashpot. Those fond of acronyms
will shortly be treated to an example of PLS, for partially latching spring.
194 6. Numerical Simulations

(a) (b)

~o
u.
Ratchet "'iii
Mechanism E
o
z

Normal Elongation

FIGURE 126. Mechanical model of the phenomenon of plasticity. It uses a set of coupled
springs, one of which activates a ratchet mechanism.

plastic deformation. As more and more notches get engaged, the operating point
describes increasingly skinny triangular sectors, as shown in Figure 126(b). The
plastic limit corresponds to the first notch on the ratchet. As long as that condition
is not exceeded, the regime remains linear with a stiffness K j. Beyond that point,
two offset springs act in parallel, with a net stiffness K j - K 2 . As intriguing as
this device may be, it still cannot reproduce Hertz's nonlinear penetration regime.

6.3.2. MD Collision Model


This section deals with the equations that govern the collision between two par-
ticles. We begin with a linear elastic model, and move on next to discuss the
nonlinear elastic regime?
Linear Model of a Binary Collision
We start with the simple case of two spheres colliding head-on, i.e., along a line
joining their centers. If the distance between the surfaces of two particles of index
i and j is denoted x, the differential equation governing x reads

mi mj

where fei) = fe~) + f//), since only the normal force comes into play in a head-
on collision. For simplicity, the vector notation has been dropped. The previous
equation applies only when x = ~ (d; + d j ) - rij > O. Under these conditions, we
have
d 2x dx 2
- 2 +~-+wox=O, (6-9)
dt dt
7 Soas to limit ourselves to these regimes, we will refrain from discussing the simulations done by
Taguchi, who added to the equations a viscous dissipative term [105], [106].
6.3 Molecular Dynamics (MD) Simulations 195

where ft is a coefficient describing a dissipative term introduced earlier in (6-7).


Here we have ft = D n and Wo = J K / mij, where, as usual, mij is the reduced
mass of the system of two particles. We immediately recognize the equation of a
dampened harmonic oscillator, whose well-known solution is

x(t) = ~o e- Mt sin(iVt) , (6-10)


w
where Vo is the relative velocity just before the collision, and iV is the frequency of
the dampened oscillation, with iV = JW5 -
ft2. The rate at which the distance x
varies is given by

dx Vo -Mt [ . (_) _ (_)]


dt = iVe -ftsmwt +wcoswt . (6-11)

The duration tc of the contact is provided by the expression


n: n:
tc =- = .
iv J(K/m) - (D/m)2

Contact ends when x(tc) becomes negative. Note that in the present model, tc is
independent of the relative velocity of the particles. We may define the equivalent
of the coefficient of restitution E: introduced in Section 2.2.2 by writing
[dx / dt]t=t,
E == - ,
[dx/dt]t=o
which leads to

This last relation clearly demonstrates the link between the loss of momentum
during collision and the dissipative term D n (or ft). The coefficient of restitution
also turns out to be independent of the relative velocity.
We are now in a position to calculate the maximum penetration depth X max along
the same line we followed for Hertz's model (Section 2.2.2). Maximum penetration
is obtained when the penetration velocity dx/dt vanishes at time t = tmax ' From
(6-10) and (6-11), the result is

X max = Vo _"~ t mox SIn


--;:;-e . (_
wtmax ) = Voe _ ~"lfiJ SIn
- iv ).
. -1 ( -
w Wo Wo

If the system is only slightly dissipative (for instance, when E: ::: 0.9), then wo» ft,
and tmax approaches the value 2tc , as in the case of Hertz's model. Under these
circumstances, X max reduces to
Vo
X max =-.
Wo
In other words, the penetration depth is then proportional to the relative velocity of
the colliding particles. This result differs significantly from Hertz's model, which
196 6. Numerical Simulations

a
predicts a much weaker dependence (as v I/5 ). We thus arrive at the conclusion
that a linear elastic model deviates substantially from the physical behavior of
real collisions. 8 It seems necessary to devise a more realistic model incorporating
the nonlinear nature of the contact interactions. That is the objective of the next
paragraph.

Nonlinear Model of a Binary Collision


Using the same notation as previously, we write a generalized form of the differ-
ential equation (6-9)

which may be rewritten in a more standard form [60]

d x
2 (X)Y dx (X)fJ
m-+1]d - -+Ed - x=O, (6-12)
dt 2 d dt f3
where E depends on Young's modulus and Poisson's coefficient of the material,
and 1] depends on the compression as well as the viscosity with respect to shearing.
We note in passing that the dissipative term in this last equation corresponds to a
purely viscoelastic interaction. As such, the equation does not account for plastic
deformations, permanent distortions, or dissipation of vibrational excitations via
phonons, all of which were mentioned during our discussion of Hertz's model in
Section 2.2.2.
It is informative to consider a few particular cases in terms of values of the
exponents f3 and y:
(1) f3 = 0 and y = 0 corresponds to the linear interaction described by (6-9).
(2) f3 = ~ and y = 0 corresponds to the situation described by Hertz's equation.
This can be verified as an exercise.
(3) f3 = ~ and y = ~ corresponds to a generalized situation (Kuwabara and
Kono model) in which a viscoelastic compression is added to the normal
elastic interaction [107]. In this model, the nonlinearity stems from purely
geometrical properties of the penetration.
It should be fairly evident by now that modeling collisions between particles is not
easy. The physics of contact interactions is inherently complex and remains poorly
elucidated. Furthermore, good numerical algorithms are tricky to develop, because
they have to scrupulously take into account all the time constants involved (such
as the duration of collisions, the relative velocities, the time of free flights, and
others). Carelessness is likely to lead to unphysical results. To illustrate the point,
we proceed to discuss a completely artificial effect that comes up in models based

8We might come to the elToneous conclusion that the present simple model, based on coupled spring
and dampener, is useless. In fact, it can be shown that, as long as the contact duration tc is chosen
judiciously, in other words, realistically from the standpoint of the physics of the materials involved,
MD simulations yield results that turn out to be fairly satisfactory.
6.3 Molecular Dynamics (MD) Simulations 197

on soft spheres [60]. It has come to be known as the "detachment effect," because
it causes an unphysical separation of particles undergoing multiple collisions. It
can best be understood by pursuing the simple model used in Section 3.2.1 to
describe the behavior of a one-dimensional stack of spheres subjected to a vertical
sinusoidal excitation.

The Detachment Effect


This effect comes up in both one- and multidimensional configurations. It results
from certain limitations of numerical simulations in hard-to-treat cases when the
separation between spheres is comparable to the penetration depth. In light of
our comments about the LRV procedure (Section 6.2.2), it is not hard to predict
that such situations may lead to erroneous numerical predictions. To highlight the
difficulties involved, we define an effective coefficient of restitution Ceff, consistent
with Section 2.2.2, by means of the expression

ceff=&,

where Eo and E f are the initial and final kinetic energies (before and after the
collision). It is important to choose a suitable variable to analyze this problem.
Numerical simulations suggest that one such variable is the ratio cr = so/(votc ),
where So is the initial distance separating the colliding particles. Figure 127 shows
how ceff depends on cr. The trend indicated in the figure seems to be "universal" in

0.5
ED-LRV

0.0
-6 -4 -2 o 2

log(a)
FIGURE 127. Effective restitution coefficient 8 plotted against the ratio SO/vote (see text).
The horizontal line at 8eff "'" 0.34 corresponds to the result of the ED-LRV procedure
described in Section 6.2.2. The present results were obtained for a column of ten spheres
and with fixed walls. The parameters used for the calculation were: d = 1 mm, 8 = 0.9
(the true coefficient of restitution), tc = 0.0022 s, and va = 0.03 m/s (after [59]).
198 6. Numerical Simulations

the sense that the results of MD simulations obtained for an extremely wide range
of values of tc (over three decades) and Va (varying by a factor of 400) all line up
along the same curve.
The graph reveals a sudden change in the behavior of a column of spheres when
CJ = 1, that is to say, when the separation between particles becomes comparable to
the distance traveled during the duration of a collision. As the separation becomes
smaller than this critical value, the effective coefficient of restitution approaches
unity. Such a result is in flagrant contradiction not only with experiments, but
also with theoretical predictions that 8 eff should be a decreasing function of the
number of particles. Here we find, instead, that it becomes equal to or larger than
the coefficient of a single sphere, for which 8 = 0.9. Put another way, the column
appears far too "elastic," which from a practical point of view leads to an artificial
separation of the colliding particles. No such problem exists when the initial sep-
aration is sufficiently large (CJ > 1), in which case the molecular dynamics model
agrees quite well with the results of the ED method. The latter technique, coupled
with an LRV procedure, correctly predicts that 8 eff does not depend on CJ. The arti-
ficial decompaction just discussed is at the origin of the designation "detachment
effect," whose meaning is further illustrated in Figure 128.
Here the effect is clearly demonstrated when the particles are initially in contact.
If we were to repeat the same calculations with particles initially separated by about
0.01 mm, the two techniques would produce virtually identical results.
There is, incidentally, another related phenomenon, known as the brake failure
effect, when particles collide tangentially [108]. It comes about for very much
the same reason. Here again, particles are slowed down considerably less in MD
simulations than in other, more realistic, mechanics-based models.
We conclude this brief review of molecular-dynamics models with a more
general remark, which in fact applies to all other simulation techniques as well.

0.2

ED

E
oS 0.1
~
0
t5Q)
.~

I-

0.0

0.25 0.5 0.75 0.25 0.5 0.75


Time (ms) Time (ms)

FIGURE 128. Trajectories of the centers of ten spherical particles. The MD model was
carried out using exponents f3 = ~ and y = 0 (Hertz's model). Other parameters were:
6
E: = 0.86, and t c = 6 x 10- s for a binary collision, va = -0.2 mis, and Sa = O. The ED
model used the same values (after [59]).
6.4 Simulation of the Dynamics of Contacts 199

The results generally converge as long as the colliding particles spend most of
their time sufficiently far apart, in which case the dynamic behavior of the entire
system can be accurately modeled by a series of binary collisions. As soon as
more than two particles come in contact at the same time, several questions come
up. Are the collisions binary or ternary, or worse? Are we dealing with blocks?
The answers are never simple, even from a straight physics point of view. All
simulation techniques pay a price for this fundamental indeterminacy, although
the symptoms may differ in each case. These difficulties manifest themselves in
the form of inelastic collapse in ED models, or the detachment effect in their MD
counterparts. As we pointed out in Section 3.1.1, short of knowing the details
of interactions on a microscopic scale, we find ourselves rather helpless when it
comes to predicting the dynamical behavior of a simple stack of as few as three
particles when they are almost in contact.

6.4. Simulation of the Dynamics of Contacts


Fueled by a number of advances and remarkable successes, this technique is cur-
rently enjoying increasing popularity [109], [110]. It is rooted in basic research
work on the mechanical properties of contacts. As we have emphasized repeat-
edly, the physics of granular materials is essentially governed by the mechanical
properties of contacts. The merit of the technique we are about to present is to
incorporate, as accurately as possible, a description of the various interactions
between solids, consistent with the picture developed in Chapter 2. ED and MD
methods are inherently dynamic in nature, and as such, they are ill-equipped to
deal with prolonged contacts; indeed, they are essentially useless for modeling the
static properties of granular piles. This points to the need for improved models. As
noted in previous chapters, solid friction introduces not only an indeterminancy in
the forces of contact (Section 3.1.1), but also complex stick-slip phenomena, all
of which can be traced to the discontinuous nature of the forces involved when two
solids in contact are displaced tangentially. By all indications, these discontinuities
should playa critical role in the dynamic properties of granular materials.
Unfortunately, precisely because these forces are discontinuous, it is virtually
impossible to write down an expression of the type T = f (VI' Yt) that describes
how the tangential friction force depends on the velocity and acceleration of the
objects in contact (VI and YI are the tangential velocities and accelerations, re-
spectively). That is, of course, a major obstacle to devising an exact numerical
treatment of the problem. Faced with this challenge, a number of researchers have
proposed ways to "tame" the laws of contact, as it were. The underlying idea is
illustrated in Figure 129.
It is useful to examine in more detail the nature of these irregularities. We
distinguish three possible situations:

• As we know, the law of dry friction exhibits a discontinuity-more precisely,


an indeterminacy if we do not know the past history of the contact-when
200 6. Numerical Simulations

T N
/-!N

o o
v D

o
v D

FIGURE 129. The diagrams on the left correspond to Coulomb's law of dry friction. T and
N are tangential and normal forces. The diagrams on the right show Signorini's conditions.
D is the distance between contact points. Both upper diagrams are discontinuous. The dis-
continuities have been partly mitigated in the lower figures. The gentler form of Coulomb's
law implies a viscous interaction in the vicinity of the contact, while that corresponding to
Signorini's condition assumes an elastic reaction when the solids get close to each other
(after [Ill]).

the tangential velocity at the point of contact is zero. In this case, Vt =


and Yt = 0, and the tangential resistance force can take on any value between
°
- fl.-s Nand +fl.-sN. The contact forces are not activated. Static friction exactly
offsets all other forces applied to the contact, preserving the condition Yt = 0.
• If a sufficiently large tangential force is applied, the contact gives way (Yt f=
0). The equality T = -fl.-sN sign (Yt) holds, even just before motion actually
begins (i.e., when we still have Vt = 0). In this situation, the forces of contact
are fully activated.
• The contact is said to be gliding when V t f= 0, in which case T = - fl.-d N sign
(Vt).

A similar analysis can be done on the basis of Signorini's conditions, which deal
with the normal, rather than tangential, forces. They apply to hard objects, consid-
ered impenetrable in the sense defined earlier:

• When V n = Yn = 0, the normal force opposing penetration can have any


value N :::: 0.
°
• Contact is broken the moment V n = and Yn > 0, in which case the normal
force N must vanish.

From this point of view, Signorini's conditions exhibit very much the same type
of discontinuity as Coulomb's law.
6.4 Simulation of the Dynamics of Contacts 20 I

Static

T Dynamic T
!-I,N
!-IN !-Id N

o 0
v V

-!-IdN
-!-I,N

FIGURE 130. Schematic interpretation of the indeterminacies concerning the friction


forces (after [Ill]).

An interesting exercise is to try to rederive the indeterminacies mentioned earlier


on the basis of the above diagrams. 9 The fundamental equations of the dynamics
of a system of two particles of reduced mass mred in contact can be written (in
projection on the principal axes at the contact point) as lO

and

where <t>t and <t>n are the normal and tangential components of the reaction force
due to friction. These components depend on the mode of contact between the
two particles, but not on the external forces, since we deliberately treat the two
separately. If we work in the frame of reference attached to the contact point
between the two particles, the fundamental equations are represented in the above
diagrams by straight lines with a positive slope. These straight lines would intersect
the discontinuous curves at a single point, as shown in Figure 130, which implies
a unique solution. The problem is somewhat more complicated in the case of
dry friction. Whether the solution is unique or not depends on the experimental
conditions and the way Coulomb's friction is modeled. Figure 130 reveals the
following:

• If dry friction is modeled with a single coefficient fh = fhs = fhd, the solution
is always unique for a dynamic interaction.
• For a static interaction, the straight line describing the fundamental equation
becomes vertical, and the solution becomes undetermined (with an infinity of
solutions).

9 An excellent analysis of these indeterminacies and how to handle them mathematically can be found
in [111].
lOHere we neglect any possible rotation of the particles. It could easily be added to the equations, but
it would not materially change the argument.
202 6. Numerical Simulations

• If dry friction is modeled with two different friction coefficients /-is and fLd,
with fLs > /-id, the straight line can intersect Coulomb's graph at two distinct
points, and the solution is obviously not unique. Which solution the system
chooses depends on its prior history, which opens the door to the kinds of
hysteresis effects discussed in Sections 2.3 and 3.1.1. That is a commonly
recognized characteristic studied in structural analysis [112].

These considerations may well elicit growing skepticism that it will ever be possible
to accurately model any granular system that is subject to such intrinsic indetermi-
nacies. As we now know, these indeterminacies all come from the discontinuous
character of the static resistance force. One way to get around this problem is to
resort to well-behaved functional dependences of the type depicted in the lower
part of Figure 129. Another way is to consider the static situation simply as a
limiting case of the dynamic problem (when v -+ 0). Such arguments may indeed
be viewed as a posteriori justifications of the MD and ED simulations techniques
which, being inherently dynamical approaches, avoid these problems entirely. It
is also essential to bear in mind that we have considered only hard objects (in the
sense of the hard spheres in ED simulations). The creation and rupture of micro-
contacts is unlikely to be as discontinuous as implied by the standard constitutive
laws. It is in fact quite plausible that smoother functions might describe real phe-
nomena more realistically. What we can say with some confidence is that various
models based on the arguments presented above generally lead to results in good
agreement with experiments [Ill]. This includes ED and MD simulations, as well
as others to be discussed in the latter part of this chapter. There is no compelling
reason to promote anyone technique over another. In all likelihood, a particular
approach, based on specific simplifying assumptions, can be perfectly adequate in
certain circumstances, and completely break down in others. The best strategy is
to be flexible and keep an open mind.
We proceed next to discuss two more simulation techniques, based on pro-
cedures for synthesizing piles. These methods may appear somewhat primitive
when compared to the ones reviewed thus far. Yet, they too turn out to produce
very satisfactory results, at least when the geometry of a pile is an important factor.

6.5. Monte Carlo (MC) Simulations


There exists an extensive literature on the topic of Monte Carlo simulations. It is
not our purpose here to offer a comprehensive analysis of the technique, which has
been used to solve a great many problems in statistical mechanics, among other
applications. Instead, we will highlight its ability to provide numerical solutions
to some important problems in the physics of granular materials. We will do so
by using the celebrated example of the "Brazil nut problem" [88], [92], [113]. As
discussed in Section 5.2, it involves the phenomenon of segregation by size. We
will later on introduce a rather different approach, known as the method ofsteepest
descent, which has also proven very useful.
6.5 Monte Carlo (MC) Simulations 203

We start by emphasizing the sequential character of the stacking method used by


both methods. Time is managed in both cases by providing for a relaxation phase
between successive stages of building up the stack. In that sense, we are dealing
here with a truly sequential procedure of the type we described in connection with
the SOC cellular automation model (Section 4.2.1). The chain of events can be
represented symbolically by

preparation:::} relaxation:::} stacking:::} relaxation:::} etc.

Let T be one period of the stacking-:-relaxation cycle. It is worthwhile to pause a mo-


ment to understand the implications and limitations of this strategy, in the light of
what we have learned in previous chapters about collision models and the behavior
of granular piles. First of all, this procedure obviously overlooks the dynamical
properties of collisions. Barring additional refinements, it ignores all the problems
associated with solid frictional dissipation, whether static or dynamic. Accord-
ingly, we should not expect this approach to properly describe the behavior of a
collection of particles in frequent collisions. To be more specific, we designate by
TI the time interval between the two closest sequential events defining the dynam-
ics of the pile. In the language of ED modeling of a vibrated one-dimensional stack
(Section 3.2.1), TI would be the time between two successive collisions. As we
have seen, this time can become infinitesimally small, giving rise to what is known
as "inelastic collapse." Under these same circumstances, tracking the evolution of
the system with an MD method would require sampling with a period T < TI,
which could easily entail prohibitive computational times. Short of that, the subtle
details of the mechanics of systems undergoing multiple collisions would be at risk
of being missed. This would be equivalent to neglecting events on a short spatial
scale A (of the order of the distance separating particles), which could lead to er-
roneous results. The Monte Carlo method specifically deals with successive states
of a granular medium after it has relaxed. As such, it is particularly well suited to
describing the physics of granular objects over fairly long time intervals, such as
when they are excited periodically and sufficiently slowly to leave enough time for
the pile to relax between successive excitations. I I With these precautions in mind,
the stacking techniques discussed here can be extremely valuable, notably for the
purpose of analyzing the phenomenon of segregation by size [113]. The next para-
graph outlines the practical steps required to implement a Monte Carlo simulation.

Monte Carlo Technique for Stacking and Relaxation


The methodology described here was originally used in numerical simulations
of the Brazil nut problem [88], [113]. It subsequently benefited from a number
of improvements which led to results in perfect agreement with the topological
models discussed in Section 5.2.1 [92]. For pedagogical purposes, we will begin

might be worthwhile to reread the portion of Section 3.2.1 dealing with the excitation period T in
11 It
relation to the relaxation time T of the system. We can also appreciate that the Me method should
be applied preferably to materials with a low coefficient of elastic restitution 8, simply because the
relaxation following excitation is typically a fairly rapid phenomenon.
204 6. Numerical Simulations

by following fairly closely the traditional way of using Monte Carlo calculations.
We will subsequently discuss the specifics of applying the technique to granular
materials.
Although we could, without unduly complicating the problem, treat the case of
a three-dimensional pile of dissimilar granules, we consider, instead, a collection
of identical disks of diameter d. These disks, assumed to be impenetrable, are
initially arranged randomly in a hypothetical vertical two-dimensional container
without walls. 12 In practice, this is approximated by using a ring-shaped container.
The initial configuration of such a system of N disks is described by a generalized
vector encompassing the coordinates of all centers

The potential energy E g (r) of the system is given by


N
E g (1) = L mgZj, (6-13)
j=l

where m is the mass of an individual disk and Z j is the altitude of its center.
The Monte Carlo method is based on analyzing the probability P of different
configurations r, each of which has an energy E g (1). Thermodynamics tells us that

P[E g (1)] = -1 exp [E


--g(1)]
- ,
Q kT
where Q is the partition function of the system, and T is its absolute temperature.
Note that this last expression characterizes all the configurations that are equivalent
from an energy point of view, in equilibrium at temperature T. They only differ
by the actual positions of the individual disks.
The technique consists in examining the probabilities of all possible configura-
tions arrived at by moving every disk in the population within a small region of
area 82 . We write this process as a set of equations

(6-l4a)

and

(6-l4b)

where ~x and ~z are independent random variables equally distributed in the in-
terval [-1, + 1], and 8 > O. So as to ensure that the disks do not penetrate each
other during the successive trials, we require that the interaction between adjacent
particles be governed by a potential energy U (s) of a pair such that

U(s) = 0 if s?:. d (6-l5a)

12This is a crucial restriction. We have seen in Chapters 3 and 5 that walls induce convection effects
in granular media. By getting rid of them, convection is conveniently eliminated. All that is then left
are geometrical phenomena, such as "arch effects" of the type described in Section 5.2.1.
6.5 Monte Carlo (MC) Simulations 205

and

U(s) = 00 if s < d. (6-15b)

The trials conducted according to (6-13) through (6-15) must be evaluated for
plausibility against the following criteria:

• If the quantity

t"E = E(I") - E(I') :::: 0, (6-16)

the new configuration has a lower energy than the one we started with. It is
therefore retained for subsequent calculations.
• If t"E > 0, the solution I' is not necessarily rejected, as it may well be
accessible via simple thermal agitation. It is therefore assigned a probability
given by

P t"E - P[E(I")] e x
(t"E)
( ) - P[E(i')] P -kT'
- (6-17)

In tum, this probability is compared to a random number ~ uniformly dis-


tributed between 0 and 1. If P (t"E) ::: ~, the solution is retained. If not, it is
discarded, and another one is tried.

The procedure calls for jiggling every single particle of the system, until each
member of the population has had its tum, which completes one iteration. The
new configuration is then used as a fresh starting point for the next iteration, in
which all particles are moved about all over again and allowed to relax, and so
forth.
Some comments on the temperature of the system are in order. The method
we have just described is essentially what is used traditionally for Monte Carlo
simulations of Brownian systems. When it is applied to macroscopic objects like
granular materials, the significance of equations involving the thermal energy kT
raises some legitimate questions.
As pointed out in Chapter 1, the Brownian motion of typical systems of inter-
est here is entirely negligible, the ratio mg t"z/ kT being of the order of 10 12 at
ordinary temperatures. If so, (6-17), and the criterion associated with it, gives a
probability that is always practically zero. In other words, the only really relevant
equation for a granular system is (6-16), which means that the potential energy
can only decrease at each step of the iteration. This is equivalent to assuming that
the temperature of the system is at absolute zero. The clear implication is that the
system traps particles in potential wells from which they cannot escape without
collisions on a microscopic level. In accordance with our earlier discussions, it
is clear that this simulation strategy amounts to neglecting the short-range inter-
actions normally associated with multiple collisions-which are equivalent to a
local temperature of the granular. Rather, it deals fundamentally with systems in
their relaxed states.
206 6. Numerical Simulations

Despite these restrictions, this type ofsimulation technique has proven extremely
useful to model a number of situations, such as a pile of dissimilar particles. Byway
of summary, we emphasize again that, by its very nature, this particular technique
is not a good choice to describe nonrelaxed configurations, where particles spend
only a fraction of their time in actual contact. One important example is that of
fluidized beds, which are more suitably treated by ED or MD simulations.

6.6. Sequential Model of a Pile


Monte Carlo simulations were based on keeping track of the energy of the various
geometrical configurations a pile of N particles can find itself in. The system
evolved from an initial state of energy E(n to a final state of energy E(l"), with
E (I") < E (n, without us having to worry at all about the details of the relaxation
process. Another strategy would be to mimic as realistically as possible the local
mechanical properties of the system. That is precisely the idea behind the so-called
method of steepest descent [91].
The principle of the technique is illustrated in Figure 131. The objective is to
model the way particles fall down on top of each other. The algorithm can be
summarized as follows:
• Spherical particles are dropped sequentially on top of the forming pile in a
random manner, in the sense that the coordinate x where they are deposited
is a random variable, as described in [114].
• After a particle is dropped at a random spot, it follows a "natural" downward
slope, along a path described as the steepest descent, until it finds a position of
local equilibrium. Such a position, marked "stop" in Figure 131, occurs when
the vertical projection of the particle's center crosses the line connecting

5
Stop

FIGURE 131. Illustration of the method of steepest descent (after [116]).


6.6 Sequential Model of a Pile 207

the centers of the two underlying particles. We note, incidentally, that this
assumes the absence of any rebound when the particle raches that favorable
spot, in accordance with earlier remarks on these stacking methods.
• Once a particle stops, it becomes permanently embedded in the pile.
Agitation can be simulated, for instance, by perturbing the entire system upward
and leaving it to relax on its own. This can be accomplished in the following
manner:
(1) A pile is first generated, by randomly depositing particles one at a time
and allowing them to relax after each addition, using the algorithm just
described.
(2) The stacked particles are numbered in ascending order starting from the
bottom.
(3) The entire pile is raised (fictitiously), and each particle is left to fall down
individually, again using the above algorithm. The process starts with the
lowest-numbered particles and gradually works its way up. To some extent,
it preserves a memory of the pile's prior configuration. 13
(4) Steps (2) and (3) are repeated many times, thereby simulating a vertical
vibration.
This type of simulation is relatively frugal in terms of computation time. It is the
technique of choice to treat cases involving large numbers of particles in three
dimensions. However, the limitations discussed in the context of the Monte Carlo
method apply here as well. Both techniques are good choices to treat a series of
relaxed states, to the specific exclusion of rapid interactions and multiple collisions
that may occur in real systems. The method of steepest descent is particularly well
suited to dealing with problems in which geometry is of paramount importance.
It has produced results in relatively good agreement with experiments. Perhaps its
greatest claim to fame is to have predicted the existence of critical diameters in the
Brazil nut problem, similar to the ones we found analytically in Section 5.2.1. 14

13We encourage the reader to refresh his or her memory by going back to the part of Chapter 3 that
describes the various modes of decompaction of a pile nnder vertical excitation, particularly in one
and two dimensions. This will provide further opportunities to reflect on the degree of realism of the
present algorithm.
14The algorithm described here was originally developed by Jullien et al. [91]. Interestingly, early
versions did not include noise, that is to say, random fluctuations of the particles's positions during
the stacking of the pile, making it entirely deterministic. One consequence was that segregation was
precluded for <P < <Pc, whereas the analytical model in Section 5.2.1 predicts merely a change in
behavior as the critical value <Pc is crossed. A noise source was subsequently added to the model,
and a more realistic behavior did indeed result from this improvement.
Bibliography

[1] S. F. Edwards and R. B. S. Oakeshott, Phys. A 157 (1989), 1080.


[2] A. Mehta and S. F. Edwards, Phys. A 157 (1989), 1091.
[3] J. Lemaitre, A. Gervois, H. Peerhossaini, D. Bideau, and J.-P. Troadec, J. Phys. D:
Appl. Phys. 23 (1990), 1396.
[4] H. J. Hermann, J. Physique II3 (1993), 427.
[5] R. L. Brown and J. C. Richards, Principles of Powder Mechanics (Oxford:
Pergamon Press, 1970).
[6] S. B. Savage, Flows of Granular Materials (Udine: Italy, 1992).
[7] B. B. Mande1brot, The Fractal Geometry ofNature (San Francisco: Freeman, 1977).
[8] X. Wu, K. J. Maloy, A. Hansen, M. Ammi, and D. Bideau, Phys. Rev. Lett. 71
(1993), 1363.
[9] R. A. Bagnold, The Physics of Blown Sand and Desert Dunes (London: Methuen,
1941).
[10] P. Y. Julien, Y. Lan, and G. Berthault, Bull. Soc. Geol. France 164 (1993), 649.
[11] C. A. Coulomb, Acad. Roy. Sci. Mem. Phys. Divers Savants 7 (1773), 343.
[12] M. Faraday, Philos. Trans. Roy. Soc. London 52 (1831), 299.
[13] W. J. W. Rankine, Philos. Trans. Roy. Soc. London 147 (1857), 9.
[14] 1. Roberts, Proc. Roy. Soc. 36 (1884), 226.
[15] H. A. Janssen, Z. Vereins Deutsch lng. 39 (35) (1895), 1045.
[16] Lord Rayleigh, Phil. Mag. Ser. 611 (61) (1906), 129.
[17] O. Reynolds, Phil. Mag. Ser. 550 (1885), 469.
[18] R. A. Bagnold, Proc. Roy. Soc. London Ser. A 225 (1954), 49.
[19] E. Guyon and J.-P. Troadec, Le Sac de billes (Paris: Odile Jacob, 1994).
[20] H. M. Jaeger and S. R. Nagel, Science 255 (1992), 1523.
[21] S. R. Nagel, Rev. Modern Phys. 64 (1992), 1523.
210 Bibliography

[22] J. Duran, "La Physique du tas de sable," (The physics of a sand pile), Revue du
Palais de la Decouverte 21 (1994),23-224.
[23] R. M. Nedderman, Statics and Kinematics ofGranular Materials (Cambridge, UK:
Cambridge University Press, 1992).
[24] D. Bideau and J. Dodds, eds., Physics of Granular Media (Commack, NY: Nova
Science, 1991).
[25] H. J. Heimann and S. Roux, eds., Statistical Modelsfor the Fracture ofDisordered
Media (New York: Elsevier, 1990).
[26] K. K. Bardhan, B. K. Chakrabarti, and A. Hansen, eds., Non-Linearity and Break-
down in Soft Condensed Matter (Berlin: Springer-Verlag, 1994).
[27] E. Guazzelli and L. Oger, eds., Mobile Particulate Systems (Dordrecht: Kluwer
Academic, 1995).
[28] S. B. Savage and D. I Jeffrey, J. Fluid Mech. 110 (1981), 255.
[29] M. W. Richman, Acta Mech. 75 (1988), 227.
[30] M. W. Richman, J. Rheo/. 33 (1989), 1293.
[31] H. K. Pak, E. Van Doorn, and R. P. Behringer, Phys. Rev. Lett. 74 (1995), 4643.
[32] D. D. Joseph, European J. Mech. B (Fluids) 9 (1990), 565.
[33] K. Hotta, K. Takeda, and K, Iinoya, Powder Technol. 10 (1974), 231.
[34] F. P. Bowden and D. Tabor, The Friction and Lubrication of Solids (Oxford:
Clarendon Press, 1950).
[35] E. Rabinowicz, Friction and Wear of Materials (New York: Wiley, 1965).
[36] W. Goldsmith, Impact: The Theory of Physical Behaviour of Colliding Solids
(London: Arnold, 1960).
[37] B. Bernu and R. Mazhigi, J. Phys. A 23 (1990), 5745.
[38] O. R. Walton and R. L. Braun, J. Rheol. 30 (1986), 949.
[39] S. Luding, Phys. Rev. E 52 (1995), 4442.
[40] L. D. Landau and E. M. Lifshitz, Mechanics (Oxford: Pergamon Press, 1976).
[41] S. F. Foerster, M. Y. Louge, H. Chang, and K. Allia, Phys. Fluids 6 (1994), 1108.
[42] J. Duffy and R. D. Mindlin, J. Appl. Mech. (ASME) 24 (1957), 585.
[43] P.-G. de Gennes, private communication.
[44] S. N. Domenico, Geophysics 42 (1977), 1339.
[45] J. D. Goddard, Proc. Roy. Soc. (London) 430 (1990), 105.
[46] F. X. Riguidel, R. Jullien, G. Ristow, A. Hansen, and D. Bideau, J. Physique I4
(1994),4973.
[47] G. G. Batrouni, S. Dippel, and L. Samson, Phys. Rev. E 53 (1996), 6496.
[48] F. Heslot, T. Baumberger, B. Perrin, B. Caroli, and C. Caroli, Phys. Rev. E (1994),
4973.
[49] J. Grindlay, Amer. J. Phys. 61 (1993), 469.
[50] E. Guyon, in Physics of Granular Media, edited by D. Bideau and J. Dodds
(New York: Nova Science, 1991).
[51] P. Dantu, in Proc. of the Fourth International Con! on Soil Mechanics and Foun-
dation Eng., vol. 1 (London: Butterworths Scientific, 1957), p. 144.
[52] S. F. Edwards and C. C. Mounfield, Phys. A 210 (1994), 290.
[53] D. Stauffer, H. J. Hermann, and S. Roux, J. Physique 48 (1987), 437.
[54] T. Travers, M. Amrni, D. Bideau, A. Gervois, J.-C. Messager, and I-P. Troadec,
Europhys. Lett. 4 (1987),329.
Bibliography 211

[55] S. Luding, E. Clement, A. Blumen, I. Rajchenbach, and J. Duran, Phys. Rev. E 49


(1994), 1634.
[56] E. Clement, S. Luding, A. Blumen, I. Rajchenbach, and J. Duran, Internat. J. Mod.
Phys. B 7 (1993), 1807.
[57] S. McNamara and W. R. Young, Phys. Fluids A 5 (1993), 34.
[58] S. McNamara and W. R. Young, Phys. Fluids A 4 (1992), 496.
[59] S. Luding, Ph.D. Thesis, University of Freiburg, Germany.
[60] S. Luding, E. Clement, A. Blumen, J. Rajchenbach, and J. Duran, Phys. Rev. E 50
(1994), R1762.
[61] E. Clement and J. Rajchenbach, Europhys. Lett. 16 (1991), 133.
[62] P. Evesque, E. Szmatula, and J.-P. Denis, Europhys. Lett. 12 (1990), 623.
[63] J. Duran, T. Mazozi, E. Clement, andJ. Rajchenbach, Phys. Rev. E 50 (1994), 3092.
[64] S. Warr, G. T. H. Jacques, and J. M. Huntley, Powder Techno/. 81 (1994), 41.
[65] S. Douady, S. Fauve, and C. Laroche, Europhys. Lett. 8 (1989), 621.
[66] P. Evesque and J. Rajchenbach, Phys. Rev. Lett. 62 (1989), 44.
[67] E. Clement, J. Duran, and J. Rajchenbach, Phys. Rev. Lett. 69 (1992), 1189.
[68] T. G. Drake, J. Geophys. Res. 95 (1990), 8681.
[69] I. Duran, T. Mazozi, S. Luding, E. Clement, and J. Rajchenbach, Phys. Rev. E 53
(1996), 1923.
[70] S. B. Savage, J. Phys. Mech. 92 (1979), 53.
[71] S. Luding, J. Duran, T. Mazozi, E. Clement, and J. Rajchenbach, J. Physique I 6
(1996), 823.
[72] F. Melo, P. Umbanhowar, and H. L. Swinney, Phys. Rev. Lett. 72 (1994),172.
[73] E. Clement, L. Vanel, J. Rajchenbach, and J. Duran, Phys. Rev. E 53 (1996), 1996.
[74] F. Melo, P. Umbanhowar, and H. L. Swinney, Phys. Rev. Lett. 75 (1995), 3838.
[75] J. Rajchenbach, Phys. Rev. Lett. 65 (1990), 2221.
[76] R. A. Bagnold, Proc. Roy. Soc. London Ser. A 295 (1966), 219.
[77] P. Bak, C. Tang, and K. Wiesenfeld, Phys. Rev. A 38 (1988), 368.
[78] S. K. Grumbacher, K. M. McEwen, D. A. Halverson, D. T. Jacobs, and J. Lindner,
Amer. J. Phys. 61 (1993), 329.
[79] J. Maddox, Nature (London) 347 (1990), 225.
[80] H. M. Jaeger, Chu-Heng Liu, and S. R. Nagel, Phys. Rev. Lett. 62 (1989), 40.
[81] G. A. Held, D. H. Solina, D. T. Keane, W. J. Haag, P. M. Horn, and G. Grinstein,
Phys. Rev. Lett. 62 (1990), 40.
[82] R. Burridge and L. Knoppoff, Bull. Seismo/. Soc. Amer. 57 (1967), 341.
[83] J. M. Carlson and J. S. Langer, Phys. Rev. Lett. 62 (1989), 2632.
[84] I.-P. Bouchaud, M. E. Cates, and P. Claudin, J. Physique I 5 (1995), 639.
[85] J.-P. Bouchaud, M. E. Cates, J. R. Prakash, and S. F. Edwards, J. Physique 14
(1994), 1383.
[86] P.-G. de Gennes, Europhys. Lett. 35 (1996), 145.
[87] J.-P. Bouchaud, in Non-Linearity and Breakdown in Soft Condensed Matter edited
by K. K. Bardhan, B. K. Chakrabarti, and A. Hansen. (Berlin: Springer-Verlag,
1994).
[88] A. Rosato, K. J. Strandburg, F. Prinz, and R. H. Swendsen, Phys. Rev. Lett. 58
(1987),1038.
[89] Y. Oyama, Bull. Inst. Phys. Chem. Res. (Tokyo), Rep. 18 (1939), 6001.
212 Bibliography

[90] F. Cantelaube, Ph.D. Thesis, University of Rennes I, France, 1995.


[91] R. Jullien, P. Meakin, and A. Pavlovitch, Phys. Rev. Lett. 69 (1992), 640.
[92] S. Dippel and S. Luding, J. Physique IS (1995), 1527.
[93] J. Duran, J. Rajchenbach, and E. Clement, Phys. Rev. Lett. 70 (1993), 2431.
[94] K. Ahmad and 1. J. Smalley, Powder Technol. 8 (1073), 69.
[95] J. B. Knight, H. M. Jaeger, and S. R. Nagel, Phys. Rev. Lett. 70 (1993), 3728.
[96] W. Cooke, S. Warr, J. M. Huntley, and R. C. Ball, Phys. Rev. E 53 (1996), 2556-
2564.
[97] J. Duran, T. Mazozi, E. Clement, and J. Rajchenbach, Phys. Rev. E 50 (1994),5138.
[98] F. Cantelaube and D. Bideau, Europhys. Lett. 30 (1995), 133.
[99] E. Clement, and J. Rajchenbach, and J. Duran, Europhys. Lett. 30 (1995), 7.
[100] B. Sapoval, M. Rosso, and J. F. Gouyet, J. Physique Lett. 7 (1985),11.
[101] B. Roseman and M. B. Donald, British Chern. Eng. 7 (1962), 10.
[102] S. B. Savage, in Disorder and Granular Media, edited by D. Bideau and A. Hansen
(Amsterdam: North-Holland, 1992).
[103] S. D. Gupta, D. V. Khakar, and S. K. Bathia, Chern. Engrg. Sci. 46 (1991), 1513.
[104] K. M. Hill and J. Kaliakos, Phys. Rev. E 49 (1994), 3610.
[105] Y-H. Taguchi, Phys. Rev. Lett. 69 (1992), 1371.
[106] Y-H. Taguchi, Phys. D 80 (1995), 61.
[107] G. Kuwabara and M. Kono, Japan J. Appl. Phys. 26 (1987), 1230.
[108] J. Schaefer and D. Wolf, Phys. Rev. E 51 (1995), 6154.
[109] I.-J. Moreau, European J. Mech. A 13 (1994), 13.
[110] M. Jean, in Mechanics ofGeomaterialIntelfaces, edited by A. P. S. Selvadurai and
M. J. Boulon (Amsterdam: Elsevier, 1995).
[111] F. Radjai, Ph.D. Thesis, University of Orsay, France, 1995.
[112] S. P. Timoshenko and J. M. Gere, Mechanics ofMaterials (New York: Van Nostrand
Reinhold, 1972).
[113] A. Rosato, K. J. Strandburg, F. Prinz, and R. H. Swendsen, Powder Techno/. 49
(1986),59.
[114] W. M. Visscher and M. Bolsterli, Nature (London) 239 (1972), 504.
[115] S. Middleman, Fundamentals of Polymer Processing (New York: McGraw-Hill,
1977).
[116] R. Jullien and P. Meakin, Nature (London) 344 (1990), 425.
Index

Aggregate, 3, 5 Convection, 7, 87, 93


Angle of movement, 48, 121 CPP, 92, 102, 179
Angle of repose, 48, 93, 119 Critical, 122, 125
Arching, 10,60, 104, 110, 161, Critical ratio, 163
204
Aspect ratio, 74 Decompaction, 27, 53, 87, 93, 98,
Attractor, 176 101
Avalanche, 16,48,53, 127 Decompaction parameter, 74, 89
Defect, 159
Bifurcation, 83, 115 Detachment, 197
Bistability, 176 Dilatancy, 23, 65
Blockage, 3, 4, 10
Brake failure, 197 Electrostatics, 24
Brazil nuts, 203 Event driven, 77, 108

CAM,128 Fluctuations, 1
Cannon ball, 54, 64, 101 Fluidization, 53, 76, 82
Chaos, 80, 85 Fluidized bed, 15
Cohesion, 5, 16 Fractal, 23, 132, 176, 179
Collision, 34, 55, 77,193 Fracture, 16, 107, 166
Compaction, 7 Fragmentation, 53, 72, 101,
Condensation, 82 104

213
214 Index

Friction, 11, 16,27,54, 115 Penetration, 42


Frustration, 30, 111 Percolation, 55, 65, 130, 178
Plug, 3,10
Hard spheres, 186
Hertz's model, 42, 195 Redirection, 17, 70
Hopper, 11 Restitution, 36, 195
Humidity, 24 Reynolds, 18,21,65
Hysteresis, 53, 56, 122, 138, 202 Rotating drum, 152, 155, 171
Rotation, 30, 109
Indeterminacy, 56, 102, 122, 199
Inelastic collapse, 79 Segregation, 2, 12, 154
Instability, 17, 111 Self-organization, 86,111,114,
Intermittent, 124, 152 128
Intruder, 159, 166 Sifting, 178
Signorini conditions, 191, 200
LRV criterion, 79, 186 Silo, 3, 70, 153
SOC, 16, 128,203
Mixing, 2, 13 Soft crust, 44, 192
Molecular dynamics, 190 Soft spheres, 186, 191
Monte Carlo, 202 Steepest descent, 207
Stick-slip, 32, 138
Normalized acceleration, 46, 77,
106,114 Temperature, 1, 136, 151,188,205
Numerical simulation, 164, 184 Thermal agitation, 108, 154

Oyama's drum, 181 Vibration, 7, 76, 136

Das könnte Ihnen auch gefallen