Vous êtes sur la page 1sur 8

Collins, I. F. & Muhunthan, B. (2003). Géotechnique 53, No.

7, 611–618

On the relationship between stress–dilatancy, anisotropy, and plastic


dissipation for granular materials
I . F. C O L L I N S  a n d B. M U H U N T H A N y

Techniques of modern thermomechanics are used to Nous utilisons les techniques de thermomécanique mod-
analyse the stress–dilatancy relation for shear deforma- erne pour analyser la relation contrainte-dilatance pour
tions of frictional, granular materials. Central to this les déformations de cisaillement de matériaux granulaires
approach is the recognition that in general, deformations frottants. Cette méthode part du principe qu’en général,
of granular materials can involve stored plastic work. It les déformations de matériaux granulaires peuvent mettre
is recognised that in shear deformations, where the shape en uvre un travail plastique emmagasiné. Il est admis
of the grains induces the dilatancy, the confining pressure que, dans les déformations de cisaillement, là où la forme
is a constraint and dissipates no energy. Nevertheless it is des grains cause la dilatance, la pression de confinemant
shown that, in such a shear deformation, the total plastic est une contrainte et ne dissipe aucune énergie. Néan-
work rate is equal to the rate of energy dissipation, and moins, nous montrons que, dans une telle déformation de
that this is given by Thurairajah’s classical observations. cisaillement, le taux de travail plastique total est égal au
It is further shown that anisotropy is necessarily induced, taux de dissipation énergétique – fait qui est établi par les
and that the stress–dilatancy relation is given by Taylor’s observations classiques de Thurairajah. Nous montrons
well-known formula. It is also demonstrated that the en outre que l’anisotropie est nécessairement induite et
extant procedure for deducing yield conditions and flow que la relation contrainte-dilatance est donnée par la
rules from specified dissipation functions is invalid. formule bien connue de Taylor. Nous démontrons égale-
ment que la procédure existante permettant de déduire
les conditions de déformation et les règles d’écoulement
KEYWORDS: friction; elasticity; plasticity; anisotropy d’après des fonctions de dissipation spécifiées est invalide.

INTRODUCTION of the plastic shear strain rate. All stresses are understood to
The term dilatancy was first coined by Osborne Reynolds be effective stresses, so that we shall not adopt the ‘prime
(Reynolds, 1885) to describe the remarkable phenomenon of notation’. Moreover he found that this result applied not
the increase in volume observed when granular materials are only when paths reached critical states but at all stages of
subjected to applied shear. This property distinguishes gran- the test paths. These important results have recently been
ular materials from most other engineering materials. A reviewed by Schofield (2000) and Muhunthan & Olcott
proper account of this dilatant behaviour must be taken into (2002). Adopting the standard notation for triaxial tests,
account when describing the stress–strain and strength be- Taylor’s work equation can be recast in the form
haviour of granular materials and soils.
^
p e_ pv þ q e_ pª ¼ Mpj e_ pª j ¼ Ö (1)
One of the earliest attempts to account for the increased
shear strength due to dilatancy in dense sand was by D. W. where e_ pv and e_ pª denote the volumetric and shear compo-
Taylor (1948). Taylor used the term interlocking to describe nents of the plastic strain rate tensor respectively (these rates
the effects of dilatancy. He calculated the power at peak stress can be interpreted as increments). The right-hand side of this
for some direct shear-box data and found that the energy input relation was interpreted as the plastic dissipation rate Ö ^ , so
is partly dissipated by a critical-state friction component and that in this model this plastic dissipation rate is always just
partly by the work needed to increase the volume. He used the M p times the magnitude of the plastic shear strain rate.
relation x – y ¼ x, where  is the applied shear stress,  Based on a minimum rate of internal work assumption,
is the constraining normal stress, x and y are the horizontal and Rowe (1962) related Reynolds dilatancy to the principal
vertical coordinates, and  is a friction coefficient. stress ratio and termed this relationship the stress–dilatancy
Thurairajah (1961) performed a number of triaxial shear relation. The stress–dilatancy relation has since become one
tests and calculated the proportion of work that went into of the fundamental conceptual ideas in geomechanics. The
Taylor’s dilation and the proportion that went into the change relation attempts to model the increase in strength that
of elastic energy in an effectively stressed soil. His work occurs when a soil dilates in response to applied shear
with both drained and undrained tests on kaolin clays and deformations.
sand led to some remarkable observations. He found that the Many of the constitutive models for soils based on
rate of work dissipated in plastic deformation is equal to the concepts of work and dissipation, interpret equation (1) and
product of the effective mean normal stress, p, with M, similar forms as a stress–dilatancy relationship. For exam-
the coefficient of friction at critical state, and the magnitude ple, equation (1) can be rewritten as
 ¼ d þ M, or tan  ¼ tan ł þ tan c (2)
Manuscript received 2 August 2002; revised manuscript accepted 4
March 2003. where
Discussion on this paper closes 1 March 2004; for further details
see p. ii.   tan   q= p, d  tan ł   e_ pv = e_ pª , and M ¼ tan c
 Department of Engineering Science, School of Engineering,
(3)
University of Auckland, New Zealand.
† Department of Civil and Environmental Engineering, Washington The angles , ł and c are the mobilised friction angle,
State University, Pullman, WA, USA. dilation angle and critical-state friction angle respectively.

611
612 COLLINS AND MUHUNTHAN
(Note here we are adopting definitions appropriate for of thermomechanics to explore the relationship between the
compressive triaxial tests. The corresponding relations for stress–dilatancy relation in the case of dilation induced by
extension tests will be discussed later.) Following on Rowe’s shearing (as envisaged by Taylor and Schofield & Wroth)
analysis of the energy expended by rods sliding over each and the plastic dissipation function. It will be shown that it
other, the above interpretation can be deduced by consider- is necessary to include anisotropic responses in order to
ing the work done in sliding two ‘interlocking saw blades’ generate dilatational strain rates, and that, when anisotropy
or ‘rough surfaces’ over each other, as described in the is included, Taylor’s relation (equation (1)) is predicted by
books by Atkinson & Bransby (1978), Bolton (1987) or the theory, as also is Thuairajah’s striking result that the
Wood (1990), for example. plastic dissipation is always given by Mpj e_ pª j. It will also be
In the family of critical-state-based models (Roscoe & seen that under these conditions all the energy is dissipated
Burland, 1968; Schofield & Wroth, 1968), the stress– and none is stored, and that Goddard’s ‘work-free reactive
dilatancy relationship was interpreted as an equation for the stresses’ occur naturally in the theory.
plastic potential. Invoking Drucker’s stability postulate, the
integrated form of the stress–dilatancy relationship was used
to generate the yield curves and to develop plastic stress–
strain models. Many of the extant plasticity-based models in PLASTICITY THEORY AND THE STRESS–DILATANCY
geomechanics do, in one form or another, incorporate these RELATION.
ideas. The stress–dilatancy relation (equation (2)) was used as a
Based on the modern theory of internally constrained basis for the original Cam clay model of Schofield & Wroth
continua Goddard & Bashir (1990) argued that Reynolds (1968), who realised that it could be interpreted as an
dilatancy represents a kinematic constraint. As stated in equation for the plastic potential g( p, q), as it can be
Goddard (1999): rewritten:
Reynolds dilatancy represents a strict kinematic coupling q @ g=@ p dq
¼ þM þM (2a)
between shape and volume for assemblies of rigid p @ g=@q dp
particles, such that, in the limit of frictionless granules which integrates to give
envisaged by Reynolds, the plastic yield locus is  
tantamount to a purely reactive (work-free) stress. pc
q ¼ Mpln (4)
p
Furthermore, he states that:
so that invoking Drucker’s stability postulate, which requires
Rowe’s stress–dilatancy concept is based on the recogni-
a normal flow rule, equation (4) could also be used as that
tion that intergranular friction results in an additional
of the yield surface, with pc being interpreted as the normal
active stress (in particular, dissipative), determined by the
consolidation pressure.
above reactive stress.
As is well known, the above model, while appropriate for
See also Goddard & Didwania (1998). These ideas are describing the plastic shearing behaviour in the neighbour-
highly pertinent to the development presented here. hood of the critical-state line, does not accurately model the
There is a vast literature pertaining to the relationship isotropic compression of a soil. Accordingly Burland (1965)
between strength and dilatancy. Here we shall only recall and Roscoe & Burland (1968) proposed a modified dissipa-
those that are particularly relevant to the present investiga- tion function Ö ^ , and replaced equation (1) by
tion. Kanatani (1982) also emphasised that it is necessary to qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
introduce internal constraints in order to understand frictional p e_ pv þ q e_ pª ¼ p e_ p2
v þ M e_ ª ¼ Ö
2 p2 ^ (5)
behaviour and dilatancy, and argued that if the constraining
stress is subtracted from the total effective stress, Drucker’s where now the volumetric plastic strain rates contribute to
postulate still applies. The concept of regarding dilatancy as a the dissipation. Equation (5) can be rewritten as a ‘stress–
kinematic constraint is also central to the theory developed dilatancy relation’:
by Houlsby (1993), who also recognised the need to introduce 2  M 2
anisotropy. Anisotropy is an essential feature of micromecha- d¼ (6)
2
nical models that use some form of fabric tensor, as in
Mehrabadi & Nemat-Nasser (1983). Related to the issue of Proceeding as above, this can be integrated to give the,
the development of anisotropy is the question of coincidence assumed coincident, potential and yield functions of modi-
or non-coincidence of the principal axes of the stress and fied Cam clay (Wood, 1990):
plastic strain rate tensors. This issue has been addressed by q 2  M 2 pð pc  pÞ ¼ 0 (7)
Oda (1975) and Gutierrez et al. (1991). Bolton (1986)
suggested a slight variation on Taylor’s and Rowe’s classical There are two important points to be made about this
results, based upon the analysis of a large number of test procedure. First, referring to equation (6) as a ‘stress–
results. Nova (1982) also proposed a minor modification to dilatancy relation’, as is done by Wood (1990) and Li &
equation (2), which is replaced by  ¼ (1  N )d + M, where Dafalias (2000) for example, is departing from the original
N is a new material parameter, and proceeded to deduce an objectives of Taylor and Schofield & Wroth. Instead, the
extended version of modified Cam clay. This model was equation models not only the volume changes induced by
further discussed by Jefferies (1997), who demonstrated that shearing, but also those due to isotropic compaction and
the extra term represented stored rather than dissipated dilation. The term is being used in a more general sense,
energy. This result is a precursor to the concepts of the and is, in essence, a statement of an elastic/plastic flow rule.
thermomechanical approach, which is central to the new Secondly, as explained later, there is a serious internal
approach employed here. Finally we note that a number of inconsistency in the logical steps leading from the work
recent papers by Li, Dafalias and co-researchers have pro- equation (5) to the yield condition (equation (7)) (and also
duced a series of models in which a state parameter is used to from equation (2) to equation (4)). Equation (5) is, in fact,
develop state–dilatancy models: see Li et al. (1999), Li & not the true work equation associated with modified Cam
Dafalias (2000), Li (2002) and Wang et al. (2002). clay. As shown by Houlsby (1981) and further discussed by
The purpose of the present paper is to apply the methods Collins & Houlsby (1997), Collins & Kelly (2002) and
STRESS–DILATANCY, ANISOTROPY AND PLASTIC DISSIPATION 613
Collins & Hilder (2002), the true work equation associated functions it has the dimensions of stress times strain rate.
with the modified Cam clay model (equation (7)) is The derivatives of Ö ^ with respect to the two strain-rate
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi components hence define two stress components,  and 
W^ p ¼ p e_ pv þ q e_ pª ¼ 1 pc e_ pv þ 1 pc e_ p2
v þ M e_ ª
2 p2 (8) say, where
2 2
The first term on the right-hand side can be positive or @Ö ^ p e_ pv @Ö ^ pM 2 e_ pª
 p ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi and   p ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
negative and represents the rate of change of the recover- @ e_ v @ e_ ª
able, stored or frozen plastic work, while the second term e_ p2
v þ M e_ ª
2 p2 e_ p2
v þ M e_ ª
2 p2

represents the part of the work rate that is being dissipated. (9)
This is an example of the general, thermomechanical decom-
position of the plastic work rate, W ^ p , into a stored part, from which it follows that
which is related to the rate of change of the free energy ^
 e_ pv þ  e_ pª ¼ Ö (10)
function, and the dissipated plastic work, which corresponds
to the production of entropy, and which must—from the This last result would hold for any dissipation function of a
second law of thermodynamics—always be positive when rate-independent material, as equation (10) is just Euler’s
plastic deformations are occurring. (Note that we are using a theorem for a homogeneous function of degree 1. There is
‘hat’ notation to denote the rate of change of plastic work clearly an inconsistency between equations (5) and (10),
and dissipation in equations (1), (5) and (8). This is because which, if both are valid, requires  ¼ p and  ¼ q, which is
these are not unique time derivatives, as plastic work and clearly not the case from equation (9). This paradox has
dissipation are path-dependent quantities.) arisen from our equating the plastic work rate W ^p
Stored plastic work has been a neglected aspect of geome-  p e_ pv þ q e_ pª to the dissipation rate Ö^ in equation (5), and
chanics. This is in no small part due to the commonly held hence neglecting the possibility of some of the plastic work
view that, in granular materials, all the plastic work is being stored.
dissipated. There is much experimental evidence leading to The stresses  and  are termed the dissipative pressure
this conclusion, such as the experimental results of and dissipative shear stress respectively, and are components
Thurairajah (1961) already mentioned, the cyclic tests on of the dissipative stress tensor @ Ö ^ =@ e_ p. They are related to
saturated silica sand reported by Okada & Nemat-Nasser the true stresses through the relations
(1994), and the infrared thermography experiments on vi-
p ¼ r þ  and q ¼  þ  (11)
brating sands by Luong (1986). However, these experiments
relate to deformations involving extensive shearing. The role where r and  are the pressure and shear components of the
of stored work in modified Cam clays and the extensions shift or back stress tensor, which define the stored work, and
discussed in the papers by Collins and co-researchers, cited are given by
above, as well as in the papers by Palmer (1967) and
@Øp @Øp
Jefferies (1997) and the book by Ulm & Coussy (2003), all r¼ p and  ¼ (12)
relate to stored work arising from isotropic compression @ev @epª
provided the model is isotropic. It is not envisaged that any where Øp , which we shall assume depends only on the
shear work is stored, unless the material is anisotropic. plastic strain components, is the plastic part of the free
These issues will be returned to and amplified below. The energy function. In general the free energy function will
model actually associated with Roscoe & Burland’s dissipa- depend on
tion function in equation (5) is the Drucker–Prager model,  both the elastic  and plastic strains and be of the
form Ø eev , eeª ; epv , epª . Here we are making the common
with a non-zero dilation angle, as shown in Collins & Hilder assumption that the material is ‘decoupled’ in the sense that
(2002) for example. thefree energy can be written as the sum of an elastic term
A possible micromechanical mechanism for producing 
Øe eev , eeª , which
 depends  only on the elastic strain, and a
stored plastic work is discussed in the papers by Collins & plastic term Øp epv , epª , which depends only on the plastic
Kelly (2002) and Collins & Hilder (2002). The essential strain. This assumption is sometimes termed the separation
point is that although, from a continuum viewpoint, an of energies hypothesis as in Ulm & Coussy (2003), and is
element is considered to be entirely plastically stressed, on equivalent to insisting that the instantaneous elastic moduli
the micro scale the stress distribution is highly inhomoge- are independent of the plastic strains. As a result the
neous. Only part of the element is at yield. In isotropic swelling lines in ev –ln p space will be parallel lines.
compression those particles in the force chains will in From equation (11) the rate of plastic work can hence be
general be plastically stressed, but many of the granular written
particles between the force chains may well be only elasti-
cally stressed. Some of this elastic energy will be recovered W^ p  p e_ p þ q e_ p ¼ r e_ p þ  e_ p  þ  e_ p þ  e_ p  (13)
v ª v ª v ª
upon unloading, but if these elastically stressed particles are
‘trapped’ within the force chain network, this elastic energy which, using equations (10) and (12), can be rewritten
can be recovered only by reversed plastic loading. Such _ pþÖ
^p ¼ Ø
W ^ , where Ö
^ >0 (14)
‘trapped’ elastic energy would provide the stored energy that
appears as stored plastic energy in the continuum viewpoint. This is a concise statement of the first and second laws of
The relationship between hardening and stored, or frozen thermodynamics for isothermal deformations applied to the
plastic energy, is discussed at length in the book by Ulm & plastic deformations of a decoupled, elastic/plastic material.
Coussy (2003). Returning to the particular dissipation function defined in
equation (5) we see by eliminating the plastic strain rates
from equation (9) that the two dissipative stresses satisfy the
DISSIPATION FUNCTIONS AND THE DEDUCTION OF equation
YIELD CONDITIONS AND FLOW RULES 2 2
In order to illustrate the internal inconsistencies in the 2
þ 2 2¼1 (15)
p M p
extant procedures of deducing yield conditions and flow
rules from a given dissipation function, consider the dissipa- which is the yield condition in terms of dissipative stress com-
tion function defined in equation (5). Like all dissipation ponents. The yield loci are concentric ellipses. Moreover,
614 COLLINS AND MUHUNTHAN
taking the ratio of the two equations in equation (9) gives models). Frictional materials are those in which, in addition
the flow rule in the form to the actual stresses, the dissipation function depends on
the consolidation pressure, pc , which is a known function of
e_ pv M 2
d  tan ł   p ¼ (16) the plastic strains, but which does not introduce any material
e_ ª  constants with the dimensions of stress. All the standard
This flow rule is always normal to the yield surface in the critical-state theories fall into this class, as do the single-
dissipative stress plane. This is always true, and is one of surface models introduced by Lade (1975). The resulting
the powerful results of the general thermomechanical theory. yield loci are homothetic (self similar) curves or surfaces.
To obtain the corresponding yield condition in the true stress The final classification is that of quasi-frictional materials,
plane, we must add on the shift stress components. For sake which, in addition to the actual stress, depend on material
of illustration we shall take r ¼ p, so that  ¼ 0 from parameters that do involve stress, such as the cohesion, as in
equation (11), and  ¼ 0, so that  ¼ q. The yield condition structured soils, or fracture toughness, as in the particle-
(equation (15)) hence becomes q ¼ M p, while equation (16) crushing models of McDowell et al. (1996), McDowell &
shows that the flow rule is incompressible and hence no Bolton (1998), McDowell (2000) and McDowell et al.
longer normal. The resulting constitutive equation is hence (2002).
that for a linear-frictional, Drucker–Prager type material Our concern here is with purely frictional models. Aspects
with an incompressible flow rule—a result that is very of the parallel theory for frictional materials are given in the
different from the normal flow rule, modified Cam clay paper by Collins & Hilder (2002). We shall consider models
model obtained by the extant procedure. This example in which the dissipation function is taken to be
illustrates the general result established by Collins & Houls- qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
by (1997), that the flow rule in true stress space is non- ^ ¼ p ( e_ pv þ tan Ł e_ pª )2 þ M 2 e_ p2
Ö ª (17)
associated, whenever the dissipation function depends on any
of the true stress variables ( p in this case). It is to be noted This is the anisotropic extension of that used by Roscoe &
that many different free energy functions can be found, Burland. It reduces to equation (5) when the ‘anisotropy
which give rise to the shift stresses r ¼ p and  ¼ 0. These angle’, Ł, is zero. This dissipation function was proposed by
correspond to different hardening behaviours and hence to Muhunthan et al. (1996) and Masad et al. (1998) and used
different evolution laws for the yield surfaces. A number of to develop yield surfaces and flow rules for anisotropic
such laws were discussed by Collins & Hilder (2002). materials. It is a special case of the general anisotropic
The yield condition and flow rule in dissipative stress dissipation function introduced by Dafalias (1986, 1987).
space are determined by the dissipation function. The yield However, these authors used the extant procedure, and their
condition and flow rule in true stress space are then obtained analyses are subject to the criticisms made above. As will be
by adding on the shift stress components. The extant proce- seen, the coupling between the volumetric and shear strain
rates introduced by the first term in the expression for Ö ^ in
dure omits the shift stress components, and invokes an
extraneous flow rule assumption, which is in general incon- equation (17) enables us to model the dilational strains
sistent with the yield condition and flow rule demanded by induced by the imposed shear strains. If, instead, we were to
the dissipation function. couple these strain rates in the last term in equation (17), we
The general theory of this new procedure, as it relates to would be able to model the shear strains induced by
geomechanics, is given in the papers by Collins & Houlsby imposed volumetric strains in an anisotropic material.
(1997) and Houlsby & Puzrin (2000). General references to The full plastic work equation can hence be written
the modern thermomechanics of continua can be found in ^ p ¼ p e_ p þ q e_ p ¼ r e_ p þ  e_ p
W v ª v ª
these references, and various applications to single-surface,
critical-state models have been developed in Collins & Kelly qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
(2002), Collins & Hilder (2002) and Collins (2003). þ p ( e_ pv þ tan Ł e_ pª )2 þ M 2 e_ p2 ª (18)
As the shear term in the dissipation function for modified p
Dividing by p e_ ª gives a general dilatancy relation, expressed
Cam clay is actually 12 pc Mj e_ pª j, as in equation (8), and not
in terms of angles:
pMj e_ pª j, as envisaged by Roscoe & Burland in equation (5),
modified Cam clay does not in fact include the fundamental (tan   tan ł) ¼ (tan Ł  tan ł)
‘Coulomb’ friction law, which is the basis of original Cam qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
clay. In consequence Collins & Kelly (2002) developed a
þ (tan Ł  tan )2 þ tan 2c (19)
new family of models in which the coefficient of M e_ pª in
the dissipation function is a linear combination of p and pc . where
As the dependence on p is increased, it was found that the
yield loci became more ‘tear-drop’ shaped, and the flow rule r 
 and tan Ł  (20)
became increasingly more non-associated. p r
The dissipative stress components corresponding to the dis-
sipation function (equation (17) are
STRESS–DILATANCY RELATIONS FOR SHEAR
@Ö ^ p2 ( e_ pv þ tan Ł e_ pª )
DEFORMATIONS OF PURELY FRICTIONAL  p ¼ ,
MATERIALS @ e_ v Ö
^
When constitutive behaviour is defined in terms of dis-
@Ö ^ p2 [( e_ pv þ tan Ł e_ pª )tan Ł þ M 2 e_ pª ]
sipation functions, it is convenient to define three classes of  p ¼ (21)
frictional materials. Purely frictional materials are those in e_ ª Ö^
which the only stress variables in the dissipation function
from which it follows that
are the stress variables themselves, typically the mean
pressure, p. Such models give rise to yield loci that are M 2 p2 e_ pª
radial straight lines in two-dimensional models, as in triaxial    tan Ł ¼ (22)
^
Ö
or plane strain situations, and to cones in three dimensions
(e.g. the Coulomb, Drucker–Prager and Matsuoka–Nakai The equation of the inclined, elliptical yield locus in dis-
STRESS–DILATANCY, ANISOTROPY AND PLASTIC DISSIPATION 615
sipative stress space (Fig. 1) is obtained by eliminating the classical technique of Lagrange multipliers. The different
plastic strain rates between equations (17), (21) and (22): approach employed here provides rather different insights.
2 As  ¼ 0, the dissipative stress is at the points A or B on
2 ð   tan ŁÞ the yield ellipse (Fig. 1). The above equations simplify
þ ¼1 (23)
p2 M 2 p2 considerably when  ¼ 0. The shift pressure, r, is equal to
the actual pressure, p, from equation (11), so that  ¼ 1 in
Inverting equation (21) and using equation (22), the normal
equation (19). The yield condition (equation (23)) and flow
flow rule for the plastic strain rates is obtained:
" # rule (equation (25)) reduce to
 ð   tan ŁÞ tan Ł ^    tan Ł  ¼ Mp and tan ł ¼ tan Ł (27)
e_ v ¼ Ö 2 
p ^
2 2
and e_ pª ¼ Ö
p M p M 2 p2 so that the dilation angle, ł, is equal to the anisotropy
(24) angle, Ł, and hence
so that the dilation angle is given by e_ pv þ tan Ł e_ pª ¼ 0 (28)
e_ pv M 2 The dissipation function (equation (17)) hence reduces to
tan ł   p ¼ tan Ł  (25)
e_ ª    tan Ł Ö
^ ¼ pMj e_ p j (29)
ª
We note that Ł is the inclination of the chord of the ellipse
which is just Thurairajah’s result. The dissipation rate is
joining points where the normal is horizontal (that is, the
always just p M times the magnitude of the plastic shear
points where e_ pª ¼ 0).
strain rate in any such constrained shear deformation. It is
very important to note that the introduction of anisotropy is
essential to the derivation of this result. If the anisotropy
STRESS–DILATANCY RELATIONS FOR CONSTRAINED angle, Ł ¼ 0, it follows from equation (28) that the
SHEAR DEFORMATIONS OF PURELY FRICTIONAL volumetric strains are also zero, and no volumetric strains
MATERIALS are possible. Again, we note that the precise form of the free
The above analysis relates to a general deformation of the energy function is not needed in this analysis, as we are
material. We now specialise to a constrained shear deforma- concerned with analysing the deformation at a given state of
tion in order to model the situation considered by Taylor hardening. Different choices of free energy functions would
and Schofield & Wroth. We shall define a constrained shear give different hardening rules and hence different evolution-
deformation to be one in which the volumetric part of the ary laws for the dilation and rotation angles, as discussed in
dissipation rate is zero: that is, Collins & Hilder (2002).
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi The dilatancy relation (equation (19)) now simplifies to
 
^ ¼ p e_ pv þ tan Ł e_ pª 2 þM 2 e_ p2
 e_ pv ¼ 0, so that  e_ pª ¼ Ö ª tan  ¼ tan Ł þ tan c (30)
(26) as illustrated in Fig. 2. It remains to determine the angle Ł
As the volumetric strain rate will be non-zero when the relating the two components of the shift stress as defined in
material dilates, this condition requires the dissipative pres- equation (20). The rate at which plastic energy is being
sure, , to be zero. Thus, while the imposed shear strain stored is
 
rates induce the volumetric deformations, these volumetric W^ s  r e_ p þ  e_ p ¼ p e_ p tan Ł  tan ł (31)
v ª ª
strain rates do not contribute to the dissipation. This is
essentially the same concept discussed by Kanatani (1982), In a parallel study of frictional materials, involving normal
Goddard & Bashir (1990) and Houlsby (1993), who regard consolidation pressures, carried out by Collins & Hilder
the induced dilatational strain rates as internal constraints. (2002), it was shown that, in order for the yield surfaces in
Houlsby modelled this situation by treating the volumetric true stress space to pass through the origin, and to touch the
strain rates as constraints on the work equation, using the

τ q Ψ

θ⫽Ψ Yield line


C.S.L.
A

π Line of centres
C
O
ζ
O p
ρ
θ

B θ

Φ ⫽ p √[epv ⫹ tanθ epγ ]2 ⫹ M2epγ 2


• • •

Fig. 2. Yield line and critical-state line for general shear


Fig. 1. Inclined elliptical yield locus. A ‘constrained shear deformation of a ‘purely frictional’ material. The centre, C, is
deformation’ occurs at A and B, where  0, and the dilation the point whose coordinates are (rr, ), the components of the
angle ł 6Ł shift stress tensor
616 COLLINS AND MUHUNTHAN
q-axis at this point, it was necessary that Ł ¼ Ł, and that shift stress is needed. This, of course, is identical to equation
this angle is the slope of the normal consolidation line. (2), as we have shown that  ¼ 1, so that r ¼ p: it follows
Hence in a ‘constrained shear deformation’ Ł ¼ Ł ¼ ł, and that /p ¼ tan ł.
so from equation (31) no plastic work is being stored. In the The effect of anisotropy and the introduction of the shear
present context we could appeal to this result and presume shift stress component is to rotate the yield loci. In a
that it stays true in the limit as the consolidation pressure, compression test these loci rotate anti-clockwise, so that in
pc , becomes infinitely large, in which case the model the extension region, where e_ pª is negative, the yield line
becomes purely frictional. moves inside the critical-state line, and the resultant volu-
Alternatively we note from equation (31) that, if Ł , ł, metric plastic strain rates are compressive. The dilation angle
the rate of change of stored work is actually negative. In the as defined in equation (3) is still positive, but the mobilised
case of a frictional or quasi-frictional material, Collins & friction and critical-state friction angles are negative: see
Hilder (2002) showed that plastic work has already been Fig. 3. Hence if the specimen is unloaded from a given
stored in the material during the pre-consolidation process, dilatational state, at constant pressure, and then sheared in
and that, in a drained test, this stored energy is reduced to the opposite direction, the specimen starts to contract plasti-
zero during the deformation prior to the onset of the fully cally, before the isotropic extensional critical-state line is
developed constrained shear deformation, which sets in when reached. As one would expect in such a kinematic hard-
the drained failure line is reached. In the present analysis of ening, anisotropic model, the material is exhibiting a
a purely frictional material, we are modelling this fully Bauschinger effect. This is also a feature of the model of
developed shear flow directly, and we have no pre-consolida- Houlsby (1993), who notes that this is entirely consistent
tion mechanism available to store plastic work. with the ‘sawtooth’ analogy, where there is a definite
The rate of change of the stored work can hence not be preferred orientation needed to produce dilation. Different
negative in such an analysis. As already noted, experimental critical-state angles in uniaxial compression and extension
observations would indicate that no work is stored in such can be built into the model by introducing some dependence
deformations, so that this rate of change cannot be positive on q in the dissipation function as shown by Collins &
either. We are hence led to the conclusion that the rate of Hilder (2002) or, preferably, by developing fully three-
change of the stored work is zero, so that once again we dimensional models, as in Collins (2003).
have Ł ¼ Ł ¼ ł. In a general deformation, volume changes occur as a
Using either argument, we conclude that in a constrained result of isotropic compaction processes as well as a result
shear deformation of a purely frictional material, the rate of of the constrained shear deformations discussed above. We
change of the stored work, W ^ s , is zero, so that all the have shown that the shear-induced plastic volumetric strain
plastic work is dissipated. Since now Ł ¼ ł, the stress– rate is equal to  tan Ł e_ pª from equation (28). It follows that,
dilatancy relation (equation (30)) reduces to the original in a general deformation, the part of the plastic volumetric
Taylor relation: strain rate that is due to compaction is equal to
tan  ¼ tan ł þ tan c (32)    
e_ pv   tan Ł e_ pª  e_ pv þ tan Ł e_ pª
It is important to note that although no work is being stored,
the two shift stress components are non-zero. The rate of This is the first term in the expression for the dissipation
working, r e_ pv , of the shift pressure is negative and exactly function in equations (17) and (18). Hence, although it does
balances the rate of working,  e_ pª , of the shear component of not appear so at first sight, this term actually represents the
the shift stress. This result demonstrates the importance of dissipation arising from the compaction processes. This re-
including the shift stresses in the analysis, even though no mark also applies to the models considered by Muhunthan et
plastic work is being stored. These results could not have al. (1996), Masad et al. (1998) and Collins & Hilder (2002).
been established using the extant procedure in which shift
stresses are ignored from the outset. None of the work done
by the constraining pressure, p, is dissipated, as a result of q Ψ
our definition of a constrained shear deformation. Instead it
produces a negative change in the stored energy—a change
Yield line
that is exactly balanced by the stored energy produced by
the shear deformation, as a result of the induced anisotropy. C.S.L.
The rest of the shear work results in energy dissipation
given by Thurairajah’s formula. The constraining pressure, p,
is hence dissipation free, which is a more exact statement of
the situation than is work free as used originally by Goddard Line of centres
(1999). C
This view is consistent with that discussed by Arthur et
al. (1991) and Dunstan et al. (1998), who proposed an ζ
O p
idealised, micromechanical model for shear deformations of ρ⬅p
granular materials, in which the existence of the force chains
is only transient. The energy temporarily stored in such a
chain is immediately dissipated, presumably owing to fric- Ψ Yield line
tional sliding, as the chain collapses and a new chain is then
formed. Hence none of the work done is permanently stored
in the network; it is all dissipated.
It is further to be noted that, although the rate of energy C.S.L.
dissipation is Ö ^ ¼ pMj e_ p j, this is equal to  e_ p and not q e_ p ,
ª ª ª
as the shear component of stored shear energy is not zero Fig. 3. Yield lines and critical state lines for a ‘constrained
in this anisotropic model. In terms of true stresses, the shear deformation’ of a ‘purely frictional’ material, where
yield condition (equation (27)) can be rewritten as r p, and ł Ł in the compressive zone and ł 2Ł, in the
ð q  Þ= p ¼ M, where again the shear component of the extensional zone
STRESS–DILATANCY, ANISOTROPY AND PLASTIC DISSIPATION 617
CONCLUSIONS Atkinson, J. H. & Bransby, P. L. (1978). The mechanics of soils.
In this paper we have demonstrated that: London: McGraw-Hill.
Bolton, M. (1986). The strength and dilatancy of sands. Géotechni-
(a) Dilatancy must be regarded as an internal kinematic que 36, No. 1, 65–78.
constraint, and cannot be derived from a flow rule Bolton, M. (1987). A guide to soil mechanics. London: Macmillan.
based on Drucker’s stability postulate. Burland, J. B. (1965). The yielding and dilation of clay. Géotechni-
(b) The extant procedures for constructing elastic/plastic que 15, 211–214.
models from dissipation functions are internally incon- Collins, I. F. (2003). A systematic procedure for constructing
sistent; they can be replaced by a thermomechanical critical state models in three dimensions International Journal
of Solids and Structures, 40, No. 17, 4379–4397.
procedure that includes the possibility of some of the Collins, I. F. & Hilder, T. (2002). A theoretical framework for
plastic work being stored, and which guarantees that the constructing elastic/plastic constitutive models for triaxial tests.
first and second laws of thermodynamics are satisfied. Int. J. Numer. Anal. Methods Geomech. 26, 1313–1347.
(c) In a ‘constrained shear deformation’ of a purely Collins, I. F. & Houlsby, G. T. (1997). Application of thermomecha-
frictional material, defined as a deformation in which nical principles to the modelling of geotechnical materials. Proc.
the applied pressure dissipates no energy, the rate of R. Soc. London Ser. A 453, 1975–2001.
dissipation is always given by Thurairajah’s formula: Collins, I. F. & Kelly, P. A. (2002). A thermomechanical analysis of
namely M p times the magnitude of the plastic shear a family of soil models. Géotechnique 52, No. 7, 507–518.
strain rate. Moreover, the dilation and friction angles Dafalias, Y. F. (1986). An anisotropic critical state clay plasticity
model. Mech. Res. Comm. 13, 341–347.
are related by Taylor’s formula, and although the shift
Dafalias, Y. F. (1987). An anisotropic critical state clay plasticity
stresses are non-zero, nevertheless there is no net build- model. In Constitutive laws for engineering materials (eds
up of stored plastic work, and of necessity the material C. S. Desai, E. Krempl, P. D. Kiousis and T. Kundu), vol. 1, pp.
becomes anisotropic under such conditions. 513–521. Amsterdam: Elsevier.
Dunstan, T., Arthur, J. R. F., Dalaili, A., Ogunbekum, O. O. &
This problem discussed in this paper again illustrates the Wong, R. K. S. (1988). Limiting mechanisms of slow dilatant
power of the thermomechanical procedure for analysing and plastic shear deformation of granular media. Nature 336, 52–54.
developing elastic/plastic constitutive laws for granular and Goddard, J. D. (1999). Granular dilatancy and plasticity of glassy
other geomaterials. It is providing new insights and models lubricants. Ind. Eng. Chem. Res. 38, 820–822.
in a variety of areas, and is capable of providing a unifying Goddard, J. D. & Bashir, Y. M. (1990). On Reynolds dilatancy. In
framework for many aspects of continuum geomechanics. Recent Developments in structured continua (eds D. DeKee and
P. N. Kaloni), vol. 2, pp. 23–35. New York: Longman.
Goddard, J. D. & Didwania, A. K. (1998). Computations of
dilatancy and yield surfaces for assemblies of rigid frictional
ACKNOWLEDGEMENTS spheres. Q. J. Mech. Appl. Math. 51, 16–44.
The authors are grateful to Joe Goddard for helpful Gutierrez, M., Ishihara, K. & Towhata, I. (1991). Noncoaxiality and
discussions on the subject matter of this paper. stress–dilatancy relations for granular materials. In Computer
methods and advances in geomechanics (eds G. Beer, J. R.
Booker and J. P. Carter), pp. 625–630. Rotterdam: Balkema.
Houlsby, G. T. (1981). A study of plasticity theories and their
NOTATION application to soils. PhD thesis, University of Cambridge.
d dilatancy coefficient Houlsby, G. T. (1993). Interpretation of dilation as a kinematic
ev volumetric strain constraint. In Modern approaches to plasticity (ed. D. Kolym-
epv , epª plastic volume and shear strains bas), pp. 119–138. New York: Elsevier.
eev , eeª elastic volume and shear strains Houlsby, G. T. & Puzrin, A. M. (2000). A thermomechanical
g plastic potential function framework for constitutive models for rate-independent dissipa-
M critical state friction coefficient tive materials. Int. J. Plasticity 16, 1017–1047.
N a material parameter Jefferies, M. (1997). Plastic work and isotropic hardening in
p effective pressure unloading. Géotechnique 47, No. 5, 1037–1042.
pc consolidation pressure Kanatani, K.-I. (1982). Mechanical foundation of the plastic defor-
q shear stress invariant mation of granular materials. Proceedings of the IUTAM con-
W^p rate of change of plastic work
^ s  Ø_ p ference on deformation and failure of granular materials, Delft,
W rate of change of stored plastic work pp. 119–127.
x horizontal displacement Lade, P. (1975). Elastoplastic stress–strain theory for cohesionless soil
y vertical displacement with curved yield surfaces. Int. J. Solids Struct. 13, 1019–1035.
r/p Li, X. S. (2002). A sand model with state-dependent dilatancy.
 normal stress component Géotechnique 52, 173–186.
 dissipative pressure Li, X. S. & Dafalias, Y. F. (2000). Dilatancy for cohesionless soils.
r shift pressure Géotechnique 50, 449–460.
 dissipative shear stress Li, X. S., Dafalias, Y. F. & Wang, Z.-L. (1999). State-dependent
 shift shear stress dilatancy in critical-state constitutive modeling of sand. Can.
 mobilized friction angle Geotech. J. 36, 599–611.
c critical state friction angle Luong, M. P. (1986). Characteristic threshold and infrared vibrother-
ł dilation angle mography of sand. Geotech. Test. J. ASTM, 9, 80–86.
Ł anisotropy angle Masad, E., Muhunthan, B. & Chameau, J. L. (1998) Stress–strain
Ł  arctan(=r) model for clays with anisotropic void ratio distribution. Int. J.
  q/p stress ratio Numer. Anal. Methods Geomech. 22, 393–416.
Ö^ rate of dissipation McDowell, G. R. (2000). A family of yield loci based on micro
Ø, Øe , Øp free energy function, elastic part of free energy, mechanics. Soils Found. 40 No. 6, 133–137.
plastic part of free energy respectively. McDowell, G. R. & Bolton, M. D. (1998). On the micromechanics
of crushable aggregates. Géotechnique 48, No. 5, 667–679.
McDowell, G. R., Bolton, M. D. & Robertson, D. (1996). The
REFERENCES fractal crushing of granular materials. J. Mech. Phys. Solids 44,
Arthur, J. R. F., Dunstan, T., Dalili, A. & Wong, R. K. S. (1991). No. 12, 2079–2102.
The initiation of flow in granular materials. Powder Technol. 65, McDowell, G. R., Nakata, Y. & Hyodo, M. (2002). On the plastic
89–101. hardening of sand. Géotechnique 52, No. 5, 349–358.
618 COLLINS AND MUHUNTHAN
Mehrabadi, M. M. & Nemat-Nasser, S. (1983). Stress, dilatancy and strain behaviour of ‘wet clay’. In Engineering plasticity (eds J.
fabric in granular materials. Mech. Mater. 2,155–161. Heyman and F. A. Leckie), pp. 535–609. Cambridge: Cam-
Muhunthan, B. & Olcott, D. (2002). Elastic energy and shear work bridge University Press.
Géotechnique 52, No. 7, 541–544. Rowe, P. W. (1962). The stress–dilatancy relation for static equili-
Muhunthan, B., Chameau, J. L. & Masad, E. (1996). Fabric brium of an assembly of particles in contact. Proc. R. Soc.
effects on the yield behavior of soils. Soils Found. 36, No. 3, London Ser. A 269, 500–527.
85–97. Schofield, A. N. (2000). Behaviour of a soil paste continuum.
Nova, R. (1982). A constitutive model for soil under monotonic and In Developments in theoretical geomechanics (eds D. W. Smith
cyclic loading. In Soil mechanics: Transient and cyclic loads and J. P. Carter), pp. 253–266. Rotterdam: Balkema.
(eds G. N. Pande and O. C. Zienkiewicz). Chichester: Wiley Schofield, A. N. & Wroth, C. P. (1968). Critical state soil mech-
pp. 343–373. anics. London: McGraw-Hill.
Oda, M. (1975). On stress–dilatancy relation of sand in simple Taylor, D. W. (1948). Fundamentals of soil mechanics. New York:
shear test. Soils Found. 15, No. 2, 17–29. Wiley.
Okada, N. & Nemat-Nasser, S. (1994). Energy dissipation in Thurairajah, A. (1961). Some properties of kaolin and of sand. PhD
inelastic flow of saturated cohesionless granular media. Géotech- thesis, Cambridge University.
nique 44, No. 1, 1–19. Ulm, F.-J. & Coussy, O. (2003). Mechanics and durability of solids,
Palmer, A. C. (1967). Stress–strain relations for clays: an energy vol. 1. NJ: Prentice Hall, Upper Saddle River, New Jersey.
theory. Géotechnique 17, 348–358. Wang, Z.-L., Dafalias, Y. F., Li, X. S. & Maksidi, F. I. (2002). State
Reynolds, O. (1885). On the dilatancy of media composed of rigid pressure index for modeling sand behavior. ASCE. J Geotech.
particles in contact. With experimental illustrations. Phil. Mag. Geoenviron. Engng 128, 511–519.
20, 469–482. Wood, D. M. (1990). Soil behaviour and critical state soil mech-
Roscoe, K. H. & Burland, J. B. (1968). On the generalized stress– anics. Cambridge: Cambridge University Press.

Vous aimerez peut-être aussi