Sunteți pe pagina 1din 14

1

Introduction

1.1 Definition and examples of granular media


From sand to cereals, from rock avalanches to interplanetary aggregates like
Saturn’s rings and the asteroid belt (Fig. 1.1), granular media form an extremely
vast family, composed of grains with very different shapes and materials, which
can span several orders of magnitude in size. However, beyond this great diversity,
all these particulate media share fundamental features. They are disordered at the
grain level but behave like a solid or a fluid at the macroscopic level, exhibiting
phenomena such as arching, avalanches and segregation.
In this book, we shall broadly define a granular medium as a collection of rigid1
macroscopic particles, whose particle size is typically larger than 100 μm (Brown
& Richards, 1970; Nedderman, 1992; Guyon & Troadec, 1994; Duran, 1997; Rao
& Nott, 2008). As we shall see in Chapter 2, this limitation in size corresponds to a
limitation in the type of interaction between the particles (Fig. 1.2). In this book, we
will focus on non-Brownian particles that interact mainly by friction and collision.
For smaller particles, of diameter between 1 μm and 100 μm, other interactions
such as van der Waals forces, humidity and air drag start to play an important role
as well. This is the domain of powders.2 Finally, for even smaller particles, those
of diameter below 1 μm, thermal agitation is no longer negligible. The world of
colloids then begins (Russel et al., 1989).
A central motivation for the study of granular media is their involvement in
many industrial and natural phenomena. It has been estimated that more than 50%
of sales in the world involve commodities produced using granular materials at
some stage, which makes granular media the second most used type of material in

1 Our definition a priori excludes very soft particles or those that fragment during the flow. However, an assembly
of liquid drops, bubbles or soft beads can sometimes be treated as a granular medium, if the confining pressure
is low enough for the particles not to deform.
2 The terms ‘granular material’ and ‘powder’ usually apply to dry grains in the air. For particles immersed in a
liquid, one speaks rather of ‘wet granular media’ or ‘suspensions’.

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
2 Introduction

Figure 1.1 Examples of granular media.

colloids powders granular media

d
1 nm 1 µm 100 µm
Figure 1.2 A classification of particulate matter as a function of the particle size:
colloids (mud), powder (flour) and granular media (a sand dune).

industry after water (Duran, 1997; Bates, 2006). Major sectors handling granular
materials include mining (extraction, transport, milling), civil engineering (con-
crete, bitumen, asphalt, embankments, ballast trains, soil stability), the chemical
industry (fuel and catalysts are often deployed in the form of grains in order to
maximize the surface of exchange), the pharmaceutical industry (from the handling

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
1.2 What makes granular matter so difficult to describe? 3

(a) (b)

Figure 1.3 Granular media are involved in many industrial and geophysical
applications. (a) The collapse of a silo. (b) Pyroclastic flow (Soufrière Hills
volcano, Montserrat Island) (photograph from Steve O’Meara, Volcano Watch
International).

of powders for the manufacture of medicine to the handling of drugs themselves),


the food industry (cereals, animal food) and the glass industry (glass is made with
sand), to name but a few. In all these areas, problems of storage (Fig. 1.3(a)), trans-
portation, flow and mixing are often encountered, which are solved by engineers
using empirical techniques.
The other major domain of application of granular materials concerns Earth
science, our soil being mainly composed of grains. From sand dunes to landslides,
from erosion patterns to pyroclastic flows (Fig. 1.3(b)), Nature offers some of
the most spectacular examples of phenomena involving granular matter. Those
phenomena are actually not limited to Earth. Dust and grains abound in space
as well, as illustrated by Martian dunes, planetary rings and ‘granular’ asteroids
(Fig. 1.1). Apart from their inherent beauty, all these natural phenomena play a
key role in shaping our environment and therefore strongly interact with human
activity. An important part of the research effort in granular media is thus devoted
to the description and prediction of natural hazards such as avalanches, landslides,
desertification and erosion of banks.

1.2 Between solid and liquid: what makes granular matter so


difficult to describe?
Despite their industrial and geophysical applications, granular media still resist
our understanding. No theoretical framework is available to describe the variety

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
4 Introduction

of behaviour observed, even in the ideal case of a medium consisting of identical


spherical particles. This situation may appear paradoxical, more than a century
after the great revolutions of modern physics! Actually, the behaviour of a single
grain is governed by the laws of mechanics, which have not changed very much
since Newton and Coulomb. Why then is the physics of a sand pile so complex?
We can try to make a list of the main difficulties encountered when dealing with
granular materials.

• Granular media are composed of a large number of particles. Let’s consider


a teaspoonfull of sugar. Assuming a grain diameter of 100 μm and a spoon
volume of 1 cm3 , the number of sugar grains in the spoon can be estimated
to be (10−2 m)3 /(10−4 m)3 , that is a million particles! This quantity is already
close to the maximum number of particles that can be simulated with present-
day computers. It therefore seems challenging to follow the movement of each
individual grain in a much larger event such as a rock avalanche or the discharge
of a silo. An alternative strategy is rather to define averaged quantities and to
model the granular medium as a continuum. One of the main issues in the physics
of granular matter is that of how to provide such a continuum description.
• Thermal fluctuations are negligible. Actually, a large number of particles is not
necessarily a serious obstacle in physics. For example, gases and liquids are
known to be well described on the macroscopic level by the Navier–Stokes
equations, while the number of molecules in a glass of water or a gas bottle
is much larger than the number of sand grains in an hourglass. However, the
key difference is thermal agitation. In liquids or gases, thermal agitation enables
molecules to rapidly reach a local equilibrium state where all possible configu-
rations are visited according to their statistical weight, enabling the derivation
of macroscopic quantities from microscopic ones. In contrast, in a granular
medium, the particles are too large to experience Brownian motion and statisti-
cal averaging over different configurations is not possible. Granular media are
thus athermal systems.3 To illustrate this, let’s compare the thermal energy and
gravitational energy of a glass bead of diameter d = 1 mm and density ρp = 2500
kg m−3 , at room temperature, T = 300 K, and under gravity, g = 9.8 m s−2 . The
thermal energy is Eth ∼ kB T = 4 × 10−21 J, where kB = 1.38 × 10−23 J K−1
is the Boltzmann constant. The gravitational energy corresponding to a ver-
tical displacement of d is Ep ∼ mgd = 10−10 J. As expected, the thermal
energy is completely negligible compared with the gravitational energy. One

3 This does not mean that temperature is always irrelevant for granular media. At the contact scale between grains,
ageing phenomena activated by temperature may occur (creep, capillary condensation, oxidation), which can
affect the overall properties of the pile, such as the angle of avalanche or the electrical properties of the medium
(see the box ‘Electrical contact between grains and the Branly effect’ in Chapter 2)

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
1.2 What makes granular matter so difficult to describe? 5

can estimate the size dc below which thermal fluctuations play a role. Taking
T = 300 K gives dc ∼ [kB T /(ρp g)]1/4  1 μm. This corresponds to the frontier
between colloids and powders given in Fig. 1.2.
• Lack of scale separation. The continuum description of granular media is also
made difficult by the lack of clear scale separation between the microscopic scale,
i.e. the grain size, and the macroscopic scale, i.e. the size of the flow. Typically,
when sand flows down on a pile, the flow thickness is about 10–20 particle
diameters. Similarly, the breakdown of a granular soil is often localized in faults,
or shear bands, of extent a few tens of grain widths. This lack of scale separation
raises questions about the validity of the continuum approach and the definition
of an elementary volume for the averaging process. In this sense, the physics of
granular media shares similarities with nanofluidics and nanomechanics, where
the effects of the size of the molecule start to play a role.
• Interactions between grains are complex. At the grain level, the laws of solid
contact between two particles involve non-trivial and highly non-linear phenom-
ena such as friction and inelastic shocks. When grains are further immersed in
a viscous fluid, hydrodynamic interactions must also be taken into account.
Those also exhibit peculiar features such as divergence at contact, due to
lubrication forces, and long-range interaction between particles in viscous
flows.
• Granular media easily dissipate energy. A bowling ball dropped into a sandbox
does not bounce. All the kinetic energy is almost instantly dissipated by collision
and friction between the grains of sand. This dissipation at the microscopic
level is an important difference from the classical systems studied in statistical
physics.
• Granular media can exhibit different states of matter. Depending on the way it is
handled, a granular material can behave like a solid, a liquid or a gas (Fig. 1.4)
(Jaeger et al., 1996). Grains can sustain stresses and create a static pile, but can
also flow like a liquid in an hourglass, or can create a gas when they are strongly
agitated. These different flow regimes can also coexist in a single configuration,
as illustrated by the flow of beads on a pile (Fig. 1.4).

This behaviour intermediate between that of a solid and that of a liquid is a


fundamental characteristic of granular media, and is shared by other disordered
materials such as foams, emulsions and pastes (Coussot & Ancey, 1999; Lar-
son, 1999). In all these systems, the medium is composed of mesoscopic ele-
ments (bubbles, drops, particles), such that thermal fluctuation is negligible com-
pared with the particles’ interaction. Above a critical density, these elements jam
and the medium behaves like a solid. To make these systems flow, one has to
apply a yield stress or decrease the density. The attempt to unify within a single

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
6 Introduction

gas

liquid

solid

Figure 1.4 Granular media can behave like a solid, a liquid or a gas, depending
on the situation.

temperature

te
sta
u id
liq
jammed state
(amorphous solids,
granular packings, stress
dense suspension,
foams)

flowing state
(grains, bubbles, drops)

1/density

Figure 1.5 A hypothetical phase diagram for the jamming transition in disordered
media proposed by Liu and Nagel (1998). The solid or ‘jammed’ state arises at
low temperature (molecular glasses), low external stresses (foams, pastes) and
high density (granular media, emulsion).

framework the physics of these jammed systems is the subject of much research.
Some authors go further and notice that there are some similarities between jammed
soft media and amorphous molecular solids (Liu & Nagel, 1998) (Fig. 1.5). By
definition, amorphous solids such as glass and rubber contrast with crystalline

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
1.3 A sketch of the book 7

solids in that they do not exhibit long-range translational order. As the tempera-
ture decreases, they do not crystallize but remain frozen in a disordered out-of-
equilibrium state, called the glassy state, which does not correspond to a global
energy minimum. This sharp slowing down of the dynamics and increase of vis-
cosity is called the glass transition, and can be interpreted in terms of an energy
landscape. For an amorphous system, the landscape is random and has many local
minima. As the temperature decreases, temperature-activated jumps between the
different potential wells become increasingly difficult and the system takes more
and more time to change configuration. The system can then be trapped in a
jammed state, pretty much like grains in a pile. This nice analogy between soft dis-
ordered media and glassy solids is the subject of much research (Berthier & Biroli,
2009; Ikeda et al., 2012). However, this general issue is beyond the scope of this
book.

1.3 A sketch of the book


The objective of this book is to provide an introduction to the different aspects
of the physics of granular media, ranging from the solid behaviour of a sandpile
to the flow of an avalanche and to very dilute media. In doing so, we could not
present a fully comprehensive and detailed treatment of every aspect of each topic.
This implies choices and a certain amount of subjectivity, which in the case of
granular media is especially true in that research in this area is still very active and
no unifying description yet exists. We hope that the many references throughout
the book will enable the interested reader to familiarize himself or herself with the
most recent work and deepen her or his understanding of specific topics.
In this book, we mainly focus on dry granular media, for which interactions
between grains are dominated by solid contact – typically sand in air. We do
not address the broad field of dilute or semi-dilute suspensions, for which interac-
tions between particles mainly occur through hydrodynamic forces (Jackson, 2000;
Guazzelli & Morris, 2012). However, the role of the surrounding fluid is crucial
and will be discussed in the later chapters of the book, when dealing with natural
phenomena such as saturated soils, submarine avalanches, sediment transport and
erosion.
The book is organized as follows. We begin with a discussion of the interaction
forces at the grain level, giving a brief overview of contact physics and hydrody-
namic forces around solid bodies (Chapter 2). We then discuss the solid regime of
granular media. Chapter 3 describes the statics of granular media (packing, force
chains, stresses) and the elastic regime of small reversible deformations (elastic
moduli, acoustics). Chapter 4 is devoted to the plasticity of granular media, when
deformations are no longer reversible. This field is historically closely related to

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
8 Introduction

soil mechanics and is also central to our understanding of the solid–liquid tran-
sition in disordered athermal media. The second part of the book is devoted to
granular flows. We begin with the case of rapid and dilute flows (the gas regime),
which benefits from the most advanced framework with the kinetic theory of
granular media (Chapter 5). We then shift to the flows of dense granular media,
when particles flow like a liquid (Chapter 6). This regime is the most commonly
observed in industry and geophysics but also the least well understood, despite
significant advances. Chapter 7 discusses the role of the presence of an interstitial
fluid between grains and is devoted to immersed granular media and very dense
suspensions. These media occur in many geotechnical and geophysical applica-
tions, since soils are often saturated with water. The last part of the book is devoted
to geophysical applications of the physics of granular media. Chapter 8 investi-
gates sediment transport and erosion. Chapter 9 gives an introduction to dynamical
geomorphology, describing gravity-driven flows and the formation of dunes and
rivers.

The natural origin of granular media


A large part of the sediments comes from the breakdown and weathering of rocks by
physical, chemical and biological action. The physical weathering is mainly due to the
formation of fractures in rocks due to dilatation, related either to temperature variations
or simply to the relaxation of the pre-stresses under which the rocks were formed
(decompression). In humid climates, cyclic freezing and thawing expand fractures due
to the variation of the volume of water. Plant roots can play the same role. Finally, rocks
can be directly eroded by the flow of water or ice, or by the collisions of grains carried
by the wind. Chemical weathering acts mainly in the presence of water and air. Some
minerals (halite, calcite) dissolve completely, and ions are removed in solution. Other
minerals such as micas and feldspars are transformed into other mineral species, which
are often of finer size (clay) and more easily entrained by erosion. These reactions
can be accelerated by biological action. Fermentation and respiration induce oxidation
of organic matter that produces water and carbon dioxide, the latter being crucial for
the reactions of dissolution (e.g. for calcite). Furthermore, microorganisms are able to
dissolve minerals by acid reaction, releasing ions (especially metallic ions).
In addition to the production by weathering, sediments are produced by volcanic
activity. Magma contains dissolved volatiles (predominantly water and carbon diox-
ide). Upon ascent-driven decompression, the solubility of these volatiles decreases,
the magma becomes supersaturated in volatiles, and gas bubbles nucleate and grow. A
magmatic foam is formed, which expands rapidly. During explosive eruptions, frag-
mentation then takes place, and the foam transforms into a suspension of ash, lava and
coarser debris. These processes give rise to soil and rock debris.

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
1.3 A sketch of the book 9

Table 1.1 The size classification of Wentworth


(1922) used in geology

Name Size (mm)

Boulders ≥256
Cobbles 64–256
Gravels 32–64
Pebbles 4–32
Granules 2–4

Very coarse sand 1–2


Coarse sand 0.5–1
Sand Medium sand 0.25–0.5
Fine sand 0.125–0.25
Very fine sand 0.0625–0.125

Coarse silt 0.0312–0.0625


Medium silt 0.0156–0.0312
Mud
Fine silt 0.0078–0.0156
Very fine silt 0.00390625–0.0078

Clay 0.0001–0.00390625

Colloid <0.0001

Geologists use a different terminology for each class of grain sizes (see Table 1.1).
This classification is more detailed than the physical classification given in the introduc-
tion, and reflects the degree of polydispersity of natural sediments. This polydispersity
depends on whether or not a sorting process occurs during the sediment transport from
active weathering zones to the sedimentation basins where they are deposited. Glacial
moraines constituted by sediments eroded and transported by glaciers are highly poly-
disperse. Fluvial deposits of pebbles, sand and silt are in comparison slightly more
sorted, and lacustrine and marine deposits even more so. Finally, aeolian deposits
and especially sand dunes are constituted by nearly monodisperse grains (due to the
so-called aeolian sieving).
The grain shape depends also on its origin. The grains of fluvial and glacial deposits
are very sharp and reflect light. During transport, each impact leaves an asymmetric
crescent-shaped trace. The grains of aeolian sand deposits are rounded, frosted and
matte. They show traces of abrasion induced by collisions with the ground, which are
characteristic of saltation.
The colour of the sand grains can be acquired during diagenesis. Reactions with
iron oxides bring about a red colour. During the expulsion of fluids, the presence of
organic matter depletes the environment in oxygen and brings about the formation of
Fe2+ ions of green colour. The quartz grains of aeolian deposits can be coated with a

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
10 Introduction

red haematitic pigment. Since grain impacts erode this rust coating, a dune is more red
when it is static and more white when it is mobile.

Granulometry
In this book, we will often model granular media as a collection of spherical particles, all
having the same size to within a few per cent. However, materials found in industry and
geophysics are in general much more complex. Granular media containing particles of
different sizes and shapes are called polydisperse, in contrast with monodisperse media
that contain only one type of grain. Granulometry is the measurement of the shape and
size distribution of a large collection of polydisperse grains (Allen, 1996).

Particle size and shape


For a particle of simple geometry such as a glass bead, size is well defined and
characterized by a single parameter, namely the diameter d of the sphere. However,
for more complex particle shapes like that of a sand grain, the notion of size is less
obvious. A common practice consists of defining an equivalent diameter of the object,
corresponding to the diameter of an ideal object that has the same geometrical properties
as the real one (e.g. volume, area). Additional parameters called shape factors are often
defined in order to improve the description of the shape, such as the ellipticity (the
length-to-width ratio) and the roundness (the ratio of the square of the perimeter
to the projected area). In practice, the choice of these parameters is highly dependent
on the method of measurement.

Distribution of particles
The particle size distribution of an assembly of Ntot grains is characterized by giving
the number N of particles having a diameter d within a range d. The probability
of finding a particle with a diameter between d and d + d is therefore N/Ntot =
f (d)d, where f (d) = (1/Ntot )(N/d) is the (normalized) particle density distribu-
tion in number. Another useful representation is the cumulative frequency distribution

in number given by F (d) = d  <d f (d  )d  , which is the probability of finding a par-
ticle with a diameter smaller than d. Figure 1.6 shows these two kinds of representation
in the case of sieved beach sand. From the distribution law, one can compute various
quantities such as the most probable diameter, the mean diameter and the distribution
width.

Measurement methods
Sieving is probably the oldest technique to sort grain, and is still in use today. The size
distribution in mass is obtained by weighing the grains retained in a stack of sieves of
different mesh sizes. The sieving technique is simple but not very precise, since small

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
1.3 A sketch of the book 11

80 100
70
ΔN 90 F×100 (%)
80
60
70
50 60
40 50
30 40
30
20
20
10 10
0 0
50 100 150 200 250 300 350 400 50 100 150 200 250 300 350 400
(a) d (µm) (b) d (µm)

Figure 1.6 The size distribution of beach sand sieved between 200 μm and
280 μm√ obtained by image analysis. In this case, the equivalent diameter is defined
as d = 4A/π , where A is the projected area of the grains. (a) A non-normalized
histogram of particle size, Ntot = 460. (b) The cumulative frequency distribution
F , expressed in percentage terms.

grains can remain trapped by larger grains. The final distribution is also sensitive to the
sieving time and to the forcing method (amplitude and frequency of vibration). More
sophisticated techniques to measure particle size distributions are now available. The
most common is light scattering. By passing grains through a laser beam and detecting
the amount and direction of the light scattered, it is possible to deduce the effective
size of particles under some assumptions (for example, for particles that are large
compared with the wavelength, light is scattered at an angle inversely proportional to
the particle size). Light-scattering methods are now widely implemented in commercial
devices (laser granulometers). They can be applied to a wide range of particle sizes
(from a few millimetres to less than 1 μm) and volume fractions (up to 40%), as
long as the medium is dilute enough for one to neglect multiple scattering. Another
way to measure a particle size distribution is to visualize the sample directly using a
microscope, which gives also information on the particle shape (Fig. 1.6(a)). There
exist also other measurement methods, such as sedimentation (the velocity of fall
of a particle in a viscous fluid is linked to its size) and acoustic techniques. In the
field, geologists commonly determine the size of large stones and boulders by just
tossing a rope over the blocks and measuring the rock diameters in contact with the
rope.

A word of caution
Characterizing precisely the size distribution of an assembly of grains is difficult and
depends on the kind of particles and on measurement methods. First, a good rep-
resentative sample must be selected from the material. For example, in the case of
measurement using a microscope and image analysis, the deposition of the particles

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
12 Introduction

on the microscope slide and the magnification of the objective both limit the range
of sizes that can be measured simultaneously. Another difficulty is that, depending
to the measurement method, size distributions are generally not obtained in terms of
numbers of particles, but rather in terms of mass, volume or length. For example, the
sieving technique gives the mass M of particles within a range of diameter d.
This corresponds to a size density distribution in mass defined by fM (d) =
(1/Mtot )(M/d), where Mtot is the total mass of the sample. The connection between
fM (d) and the number distribution f (d) seen above requires a relation between the
mass of the particles m and their diameter d. Assuming a relationship m = αρp d 3 , with
a size-independent shape factor α and particle density ρp , gives fM (d) = Kd 3f (d),

where K = 1/ d  d 3f (d  )d  is set by the normalization condition. Similar formulas
link the size distribution in length fL (d), area fS (d) and volume fV (d) to f (d), assum-
ing that one is dealing with self-similar particles: fL (d) ∝ df (d), fS (d) ∝ d 2f (d) and
fV (d) ∝ d 3f (d). One should not confuse these different kinds of distribution, which
give more or less weight to large particles (the volume of a 1-mm-diameter sphere is
1000 times greater than that of a 100-μm-diameter sphere).

A brief history of grains


Granular matter is deeply rooted in the long history of science and technology,
humankind having always handled grains in order to build shelters, secure dams or
store cereals. One of the oldest and most remarkable uses of granular materials in
architecture and construction is given by the Egyptians, who are believed to have
erected their heavier obelisks and colossi by means of giant silo chambers filled with
sand (Fig. 1.7) (Golvin & Goyon, 1987).
The study of granular media in modern science really begins with the work of
Coulomb and his article dating from 1773 entitled Sur une application des règles
de maximis et minimis à quelques problèmes de Statique, relatifs à l’Architecture
(Fig. 1.8(a)). In this article, the father of electrostatics established the foundation of
soil mechanics by investigating the stability and failure of granular heaps, using the
so-called Coulomb-wedges method.
Following this pioneering work, several scientists have been interested in the
behaviour of granular matter during their research. Worthy of mention are the stud-
ies of Chladni (1756–1827) and Faraday (1791–1867) on vibrated powders (Chladni,
1787; Faraday, 1831), the observations of Hagen (1793–1884) describing the sat-
uration of pressure with depth in silo and the dynamics of hourglasses, the work
by Darcy (1803–1858) on flows through porous media, the work by Rankine (1820–
1872) on active and passive states of pressure in granular soils, the discovery of dilatancy
by Reynolds (1842–1912) and the work of Terzaghi (1883–1963), another father of
soil mechanics, devoted to the behaviour of water-saturated granular media (Terzaghi,
1943). Among these scientists, the English engineer Bagnold (1896–1990) holds a

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
1.3 A sketch of the book 13

1. 2.

3. 4.

Figure 1.7 An illustration of the technique used by the Egyptians to erect an


obelisk by drainage of a silo filled with sand (Ancient Egypt, second millennium
BC, adapted from Golvin and Goyon (1987)). The obelisk was first dragged along
a weakly inclined ramp and stopped above a silo chamber filled with sand. The
sand was then gradually removed from the silo, allowing the obelisk to drift
downwards until it rested upright. This remarkable technique, which does not
require a pulley or a hoist, is possible thanks to the dual nature of granular matter
intermediate between liquid and solid. First, the medium is strong enough to
sustain the weight of the obelisk, but also flows when sand is removed at the base
of the silo. In addition, the pressure in static granular materials confined in a silo
quickly saturates with depth because of friction on the walls, in contrast with the
behaviour of classical liquids (we shall see this property of silos filled with grains
in Chapter 3 with the Janssen model).

special place (Fig. 1.8(b)). This ardent traveller, who was passionate about deserts and
served as an officer of the British army during World War II, established the scientific
foundations of aeolian and sediment transport in his book The Physics of Blown Sand
and Desert Dunes (Bagnold, 1941). His experiments on the rheology of concentrated
suspensions (Bagnold, 1954, 1979) also identified an important flow rule of granular
flows, which is now known as Bagnold’s law.

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002
14 Introduction

(a) (b)

Figure 1.8 (a) A painting of Charles-Augustin de Coulomb (1736–1806) by


Hippolyte Lecomte. (b) Ralph A. Bagnold in action in the desert (a photograph by
Ronald Peel taken from Bagnold’s book Sand Wind & War, University of Arizona
Press, 1990).

After World War II, the study of granular media developed in conjunction with
the expansion of soil mechanics, in particularly in the UK with the Cambridge school
and the critical-state theory. Research in granular media was also stimulated by the
prediction of natural hazards such as rock avalanches and landslides. The concept
of granular temperature, which had initially been introduced in the 1970s to describe
Saturn’s rings, yielded in the 1980s to the development of the kinetic theory of rapid and
dilute granular flows. During the same period, depth-averaged equations inspired by
shallow-water theory were introduced to describe gravity-driven flows in geophysics.
During the last 20 years, the study of granular media has kept growing, at the cross-
roads of geophysics, mechanics, statistical physics, non-linear physics and rheology.
Whether the works of current researchers will stand the test of time only the future will
tell; but they are standing on the shoulders of giants.

Downloaded from https:/www.cambridge.org/core. University of South Florida Libraries, on 03 Apr 2017 at 15:13:38, subject to the Cambridge Core terms of use,
available at https:/www.cambridge.org/core/terms. https://doi.org/10.1017/CBO9781139541008.002

S-ar putea să vă placă și