EPJ manuscript No.
(will be inserted by the editor)
Non-local rheology in dense granular flows
Revisiting the concept of fluidity
arXiv:1510.06965v1 [cond-mat.soft] 23 Oct 2015
Mehdi Bouzid, Adrien Izzet, Martin Trulsson, Eric Clément, Philippe Claudin, and Bruno Andreotti
Physique et Mécanique des Milieux Hétérogènes, UMR 7636 ESPCI – CNRS – Univ. Paris-Diderot – Univ. P.M. Curie, 10 rue
Vauquelin, 75005 Paris, France.
October 19, 2018
Abstract. The aim of this article is to discuss the concepts of non-local rheology and fluidity, recently
introduced to describe dense granular flows. We review and compare various approaches based on different
constitutive relations and choices for the fluidity parameter, focusing on the kinetic elasto-plastic model
introduced by Bocquet et al. [Phys. Rev. Lett 103, 036001 (2009)] for soft matter, and adapted for granular matter by Kamrin et al. [Phys. Rev. Lett. 108, 178301 (2012)], and the gradient expansion of the
local rheology µ(I) that we have proposed [Phys. Rev. Lett. 111, 238301 (2013)]. We emphasise that, to
discriminate between these approaches, one has to go beyond the predictions derived from linearisation
around a uniform stress profile, such as that obtained in a simple shear cell. We argue that future tests
can be based on the nature of the chosen fluidity parameter, and the related boundary conditions, as well
as the hypothesis made to derive the models and the dynamical mechanisms underlying their dynamics.
1 Introduction
Since non-locality was introduced as an interpretive framework for dense granular flows [1,2,3,4], it has become a
key concept to describe the rheology of complex fluids in
soft condensed matter. However, the connections between
the various contributions to this subject, their similarities and possible conflicts need clarification. In particular,
among the pending questions that must be answered, a
fundamental and vivid issue is the possible emergence of
non-locality as the signature of a dynamical phase transition [5,6,7,8,9]. This interrogation does not only concerns granular matter but should be apprehended in the
more general context of amorphous solids undergoing a
rigidity transition. At present, different conceptual approaches have been put forwards to describe non-locality
and several non-local constitutive relations were proposed
for granular matter. It is thus fair to ask whether these
approaches are equivalent and to which extent for example, they are similar to phase field models [10,11,12,13]
built on an underlying liquid/solid phase transition. Also
for granular matter, shear banding and apparent “creeping zones” are observed which are difficult to re-conciliate
with a simple local rheology [14], and this has been the
starting point of different propositions for non-local constitutive relations. For many, the elements of proof validating these approaches has often been a mere ”good fitting”
of the velocity profiles. A legitimate question is then to
ask whether this is sufficient to demonstrate the validity
of a particular model and moreover, what could be other
more stringent tests providing essential information on the
dynamical mechanisms responsible for non-locality.
Our aim here is to propose a critical discussion of the
concepts of non-local rheology and fluidity in dense granular matter, based on recent progresses as well as older
results. In the next section, we first review the rheology
of dense granular flows, starting from the local rheology
towards evidence for non-local effects and describing nonlocal approaches. In section 3, we discuss the concept of
fluidity. Section 4 is devoted to the differences among the
non-local constitutive relations proposed for granular matter. We end the paper in section 5 with a discussion on
further possible tests that must be performed to better
understand the mechanisms at the origin of non-local effects.
2 On the rheology of dense granular matter
2.1 Rigidity transition
When sufficiently polydisperse to avoid crystallisation, a
granular packing at rest can be considered as an amorphous solid. By definition, amorphous solids refer here to
systems that may resist to a shear stress while they do
not present any long range translational order at the microscopic scale, namely the grain size for granular matter.
Let us consider, for clarity, an ideal rheometer in which
the material is submitted to a homogeneous shear stress
σ. The system behaves mechanically as a solid if it reaches
equilibrium at a finite strain γ. It is considered as elastic
2
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
if it returns to its original state, once the stress is removed
and plastic, otherwise. Conversely, the system behaves mechanically as a liquid if it flows permanently at a finite
strain rate γ̇. The system exhibits a rigidity transition if
its dynamical behaviour switches from solid-like to liquidlike, when a control parameter crosses a threshold value.
Most soft amorphous solids present a rigidity transition
upon varying the shear stress, the threshold value being
named the yield stress σy . We hence define the yield parameter as the ratio of the shear stress to the yield stress:
σ
Y=
.
σy
(1)
We emphasise here that the existence of a yield stress is
not an intrinsic material property. It depends on the other
control parameters that are kept constant during the loading. As an example, a granular material, dry or suspended
in a fluid, displays a yield stress when the particle-borne
pressure P is imposed, whereas it does not when the volume fraction φ is imposed. This is obviously a key issue
and often a source of confusion. This question was discussed recently in the context of different amorphous particulate materials [15].
The rigidity transition is intimately related to the multistability of the energy landscape: the system has to cross
energy barriers to flow. The physical nature of the mechanisms preventing irreversible plastic deformations allows
to classify the soft amorphous materials and their corresponding rigidity transition as:
a property associated with the presence of inter-granular
friction, and the hysteresis of the effective friction coefficient has remained unexplained up to now. For grains
of mean diameter d and mass density ρp
g , the confining
pressure also sets the time-scale T = d/ P/ρg for plastic reorganizations at the granular level (microscopic time
scale). Following [30,31], one can define the rescaled strain
rate, or inertial number, as:
γ̇d
.
I ≡ γ̇T = p
P/ρg
(2)
Writing the yield parameter as
Y=
σ
σ
=
,
σy
µc P
(3)
the constitutive relation for homogeneous steady flows
takes the generic form
Y=
µ(I)
= 1 + aI n .
µc
(4)
n = 1 for grains presenting a standard friction coefficient
≃ 0.5 at contact, and n = 1/2 for frictionless grains. This
relation must be complemented by a law relating the volume fraction φ to the rescaled strain rate I:
φ − φc = −bI n ,
(5)
with the same phenomenological exponent n; a and b are
constants that depend on the microscopic details of the
– Entropic for glasses formed by thermal quenching [16,17,18];
system. This derivation, is simply based on dimensional
the rigidity transition is then called ”glass transition” analysis, in the rigid limit where the grain elasticity is ir[19].
relevant. Empirical measurements indeed show a frictional
– Enthalpic for soft elastic particles at high volume frac- behaviour with the emergence of a yield stress.
tion [20,21,22,23,24]; the free energy may result from
Relation (4) is well suited to investigate pressure-controlled
capillarity (foam, emulsion), from electrostatics, from situations. However, the same equations can entirely be rethe particle elasticity, etc. The rigidity transition is cast to handle situations where φ is fixed, and then the
then called elasto-plastic depinning transition [25,26,27]. yield stress disappears from the constitutive picture. In– Geometric for hard grains submitted to a confining verting eq. (5), one obtains:
pressure [28]; the rigidity transition is then called jamming transition [29].
ργ̇ 2 d2
P = b2/n
,
(6)
(φc − φ)2/n
Although we focus here on the last case, we will frequently
discuss connections with other complex fluids in soft mata
ργ̇ 2 d2
2/n
σ
=
µ
b
1
+
(φ
−
φ)
.
(7)
c
c
ter.
b
(φc − φ)2/n
2.2 Local rheology
Following the seminal paper by GdR MiDi [30], major improvements were obtained to provide a consistent framework to understand and model how granular matter flows.
In the rigid limit, granular matter does not present any
intrinsic energy scale and confining pressure P thus provides the only relevant scale of energy per unit volume. As
a consequence, P sets the yield stress as σy = µc P , where
µc is the ”critical” friction coefficient, which depends on
microscopic material properties (e.g. packing polydispersity, inter-granular friction, shape etc...). For a real granular material, the rigidity transition is actually subcritical,
In this representation, the shear stress clearly vanishes in
the limit γ̇ → 0, i.e. if φ < φc . From this derivation, we can
see that, even in the limit of rigid particles, i.e. without
any explicit elasticity, a granular material is compressible
under shear [32]. The pressure therefore requires time to
establish over the size of the system. Provided this timescale remains short compared to γ̇ −1 , the pressure can
be considered as defined ”instantaneously” and thus can
be used as a state variable instead of φ. This is indeed a
central assumption to neglect compressibility (Boussinesq
approximation) when using the µ(I) formulation in heterogeneous situations, : it implicitly requires that pressure
is established macroscopically over a very short time scale
and varies slowly in time and space.
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
Using this close set of constitutive relations, a quantitative agreement with numerical or experimental measurements has been reached in different configurations. In
particular, for avalanche flows of glass beads on a rough
inclined plane – an important situation as the yield parameter Y is fixed by the inclination angle – Pouliquen
[33] has derived an effective flow rule consistent with the
µ(I) rheology. This has led to a three-dimensional extension of the local rheology [34], yielding the correct scaling
laws characterising the chute flow geometry.
Dense granular suspensions have been shown to follow the same rheology, with T = ηf /P , where ηf is the
viscosity of the suspending fluid, to form the so-called viscous dimensionless number T γ̇ [35,36] – sometimes noted
J. More generally, the rheology obeyed by a granular material in a homogeneous steady state takes the very same
form as observed for soft material presenting a yield stress.
In those more general situations, the shear rate γ̇ can then
be rescaled by a plastic time-scale T , to form a dimensionless number akin to I ≡ γ̇T . This kind of rheology takes
the same generic form (4), often called a Herschel-Bulkley
constitutive relation, for all these systems.
2.3 Failure of the local rheology
Consider now a heterogeneous shear flow. Its rheology is
said local if the stress tensor at a given location is still a
function of the shear rate at the same place. Non-locality
refers to any deviation to such a local constitutive relation.
Before giving precise examples, let us discuss the choice
of such a name. It has been introduced in granular material to describe the distant transmission of momentum
through the granular skeleton, during collisions [1,2,3,4].
In the limit of rigid particles, this transport of momentum is instantaneous so that a stress, which is a flux of
momentum, can be induced by distant collisions.
In hydrodynamics, the epitome of non-locality is pressure. In a simple fluid, pressure is transmitted at the speed
of sound. At low mach numbers, the time needed for a
pressure signal to cross the entire flow is small compared
to γ̇ −1 . In this limit, pressure is determined by the incompressibility condition ∇ · v = 0, where v is the velocity
field. Taking the divergence of the Navier-Stokes equation
one therefore obtains, for a Newtonian fluid, the Laplace
equation:
∇2 P = −ρ∇ · (v · ∇v).
(8)
Pressure balances the potential part of inertial terms. The
Biot-Savart equation provides an explicit solution of this
equation under an integral form. The fact that pressure is
a function of the whole velocity field, and not only of the
local strain rate then appears explicitly. As a conclusion,
there are a priori two definitions of non-locality which are
not equivalent:
(i) Momentum is transported over large distances on a
time-scale small in comparison to γ̇ −1 and to the plastic
rearrangement timescale T .
(ii) The constitutive relation involves a second state variable, which is not a function of the strain rate, and whose
3
evolution is controlled by an independent equation, typically involving a Laplacian operator.
For instance, the kinetic theory, which is valid for dilute
and rapid granular flows, is non-local in the weak sense
(ii) since it introduces an independent field representing
the mean squared velocity fluctuations (the so-called granular temperature), which may control the stress tensor
[39]. However, the transmission of momentum remains
perfectly local in the sense (i).
Non-locality in the weak sense (ii) manifests itself through
different properties. The first one is the evidence of a
creeping flow in regions below the yield condition (Y < 1)
[30,40,41,42,43]. Instead of the expected static zone (i.e.
a solid), one observes an exponential spatial relaxation of
the shear rate γ̇. A second property is the fact that the
yield conditions are sensitive to the system size and to
the boundary conditions. In the case of granular matter,
the yield stress measured on an inclined plane depends
on the thickness of the deposit [33,30]. Also, in conditions
where the grains should flow according to the local rheology, jammed regions are identified below the flow [44].
Furthermore, for self-channelised flows close to jamming,
quantitative departure from the local rheology predictions
are explicitly shown [45]. Finally, properties pointing on
the existence of non-local effects are revealed by microrheology experiments. For example, the force-velocity relations assessed by an intruder plunged in the material
strongly depend on the presence of a distant shear flow
[46,47,48], or of a vibrating boundary [49]. Examples of
manifestation of non-locality for soft and granular matter
are given in Fig.1 and Fig.2.
2.4 Reviewing non-local models
We review here the main approaches that were put forwards to tackle these problems.
2.4.1 Cosserat approach
In the framework of plastic theories developed for soil mechanics, the formation of localised shear bands is sometimes apprehended via a Cosserat extension of the continuous elasto-plastic theory (see for example [50] and refs
inside). To describe the quasi-static state of deformations
of a granular material, new fields are introduced that couple stresses and rotations. This theory introduces a microscopic length scale describing the range of influence of the
microscopic granular rotations and provides a non-local
coupling for plastic deformations over this scale. However,
in spite of the fact that it may provide a useful regularisation technique for numerical computation methods,
its use is often seen as limited since the issue of assessing objectively the constitutive parameters and providing
consistent boundary conditions for the fields, has so far
remained a short-coming of the approach.
4
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
1.0
1.0
0.8
0.8
0.4
0.8
0.6
0.6
0
0.4
0.4
-0.4
0.2
0.2
-0.8
0.0
0.0
0.0
0.2
0.4
0.6
0.8
1.0
-10
-5
0
5
10
0
5
10
15
20
25
Fig. 1. Experimental and numerical results for soft and granular matter displaying non-locality. (a) Velocity profile
of the flow of a jammed emulsion in a micro-channel with rough surfaces, data from [5]. Local rheology would predict
a plug flow. (b) Velocity profile of a granular flow in a 2D shear cell. Local rheology would predict a linear velocity
profile. The numerical data are from [37]. (c) Experimental (triangles) and numerical (squares) velocity profiles of a
foam flow in a Couette cell, data from [38] and [72] respectively. Local theory would predict a localised failure at wall.
2.4.2 Phase field approach
As stated above, non-locality, in the weak sense, reflects
the existence of a state parameter, beyond the strain rate,
determining the stress values. As this parameter measures
how fluid the system is, following Derec et al. [51], we will
refer to it as the ”fluidity” and will note it thereafter f .
We warn the reader that in different papers, fluidity may
refer to different physical quantities. Here, we keep the
name and its conceptual definition in relation to non-local
rheology. Importantly, we consider that for our purpose,
a relevant fluidity has to be selected on physical basis,
amongst all state variables.
From a phenomenological point of view, fluidity plays
the role of an order parameter describing the dynamical transition from solid-like to fluid-like behaviour. It
was first proposed by Aranson & Tsimring to introduce a
phase field f which vanishes in the static state and which
tends to 1 in the fully fluidised state [10]. In this approach
f is therefore dimensionless. The overall shear stress is
then formally decomposed as the sum of a solid and liquid like contributions weighted respectively by 1 − f and
f . Following Landau standard derivation, the order parameter f is controlled by a reaction diffusion equation of
the form:
T f˙ = I(f ) + ℓ2 ∇2 f,
(9)
where I is a function of f parametrised by the state variables and in particular by the rescaled shear rate I. Note
that I(f ) can be designed to reflect a subcritical, hysteretic transition from solid-like to liquid-like behaviours,
as generically observed for granular matter. The microscopic time T is a characteristic time for fluidization to
occur and ℓ an elementary length scale. The diffusion coefficient of the fluidity is ℓ2 /T . The Laplacian operator results from a gradient expansion, assuming spatial isotropy.
This term is the transcription of a (weak) non-locality. As
a matter of fact, in the steady state, f is determined by
a non-linear Laplace equation, just like pressure in hydrodynamics.
This approach produces an effective rheology different from µ(I) in the sense that Pouliquen’s flow rule [33],
valid for avalanches of spherical beads, is not recovered.
However, it yields a real solid/fluid phase transition semiquantitatively close to what is observed for sandy grains,
in particular the starting and stopping heights, flow rules
and erosion/deposition waves [52,53,54,13]. Nevertheless,
it does not reproduce creep zones close to a shear band.
2.4.3 The elasto-plastic approach
The approach proposed by Kamrin et al. [55] to model
granular flows is directly adapted from the Kinetic ElastoPlastic (KEP) model introduced by Bocquet et al. [6] for
soft matter. The key concept is the fluidity, which can be
defined in an unified way as:
f=
γ̇
.
Y
(10)
We will devote below an entire sub-section to this approach.
2.4.4 Mechanically activated plastic events
An original idea to describe non-locality has been proposed by Forterre and Pouliquen [56]. It is based on an
analogy with Eyring’s transition state theory for the viscosity of liquids, where mechanical fluctuations – introduced here as a synonymous of heterogeneities – would
play the role of temperature in thermal systems. Plastic
rearrangements occur at a rate proportional to the strain
rate γ̇. They are assumed to generate at random a new
realisation of the forces on the contact network, allowing
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
5
for the formation of new weak zones where the next rear- a systematic expansion in a functional of f . The relative
fluidity κ remains finite when f goes to 0. The constitutive
rangement will occur.
At a semi-quantitative level, this approach improves relation can be expanded around the relation Y = µ(I),
significantly the local visco-plastic approach as it pro- valid in the homogeneous case, according to:
poses a physical hint for microscopic processes inducing
Y = µ(I)χ(κ) with χ(κ) ≃ 1 − κ + O(κ2 ).
(12)
non-locality for granular media. Moreover, it can predict
dependence of the stopping angle with flow height, shear
bands extension increasing with the flow rate. However, κ is positive when the point considered is surrounded by
the theoretical outcomes are more difficult to quantita- a more liquid region (higher f ). This region flows more
tively re-conciliate with the Pouliquen’s flow rule for the easily than expected from the local value of f , and the
chute flow and more importantly, it does not predict a corresponding shear stress is therefore lower. In this forthickness dependence of the avalanche starting height. In mulation, χ(κ) must thus be a decreasing function of κ,
the non-local formulation proposed by Pouliquen and Forterrewhich justifies the minus sign in front of κ in (12). Note
[56], the shear rate γ̇ obeys an integral equation, which in- that the lack of multiplicative factor in Eq. (12) defines
volves an exponential kernel, function of the stress tensor ℓ in a univocal way. Importantly, this phenomenological
and of the distance, interpreted as a Boltzmann-like fac- derivation does not depend on the nature of the mechantor. Provided the fact that for granular packing, the stress ical interaction between the shear zones; the reader may
fluctuations take place generically over few grain sizes, think of the analogy with the van der Waals gradient exthe authors assume a spatial dependence of the interac- pansion of the Helmholtz free energy at a liquid-vapour
tion kernel as a Lorentzian function decaying algebraically interface.
There are three obvious choices for the granular fluid(power −2) over this granular size.
ity.
One would be to introduce the (coarse grained) numThe relation between the shear stress and the strain
ber
of
contacts per grain Z and number of sliding contacts
rate depends non-locally on two fields i.e. on two fluidity
per
grain
ζ. In the spirit of Maxwell rigidity transition theparameters: the strain rate γ̇ and the confining pressure P .
ory,
the
fluidity
would then be defined as the distance to
The relatively fast decay of the chosen spatial kernel makes
isostaticity.
A
second
possibility is to introduce the volpossible a long wavelength expansion of the rearrangement
ume
fraction
φ,
and
to
build the fluidity as the distance
rate equation with respect to γ̇ and P . If P varies slowly at
φ
−
φ
to
the
value
φ
that
the volume fraction reaches in
c
c
the scale of the grain size, the expansion generates at the
the
limit
I
→
0.
This
poses
two
problems: first, φc depends
first order a laplacian operator, like other models. More
on
the
fraction
of
sliding
contacts;
second, one would need
precisely, the constitutive relation presents a dependence
to
introduce
the
mass
conservation
equation for φ i.e. to
2
with ∇ γ̇, suggesting that the fluidity f is the rate of
consider
explicitely
the
granular
fluid
as compressible. Fiplastic events γ̇.
nally, there is a last quantity that can play the role of the
fluidity: the inertial number I itself. It vanishes in the solid
state and increases with the degree of fluidity. Ockham’s
2.4.5 The gradient expansion of the constitutive relation
razor – law of parsimony – is obviously in favour of such a
We have ourselves followed an approach which is signif- choice, as the equations are closed without involving furicantly different from the previous ones, and which sug- ther equations, in the dense limit φ ≃ φc . This does not
gests possible candidates for the most appropriate fluidity constitute a deep scientific argument, except that simple
of dense granular flows [37,57]. Imagine the problem solved models are better testable.
The constitutive relation (12) must actually be comand a fluidity f built, which vanishes in the solid phase
and increases with the ability to flow. We then follow the plemented by another one for the volume fraction. Asstandard vision of Maxwell rigidity transition as put for- suming that φ is a local function of I, expressing the nonwards to understand jamming in granular matter. Due to local rheology with f = I with κ = ℓ2 ∇2 I/I, and with
the cooperative motion of particles along soft modes, flow- f = φc − φ with κ = ℓ2 ∇2 φ/(φc − φ) are mathematically
ing is facilitated when, at a given point it is surrounded by analogous. The only subtlety is that compressibility must
a more fluid zone. Conversely, the resistive stress is larger be taken into account if the non-local constitutive relation
when the point is surrounded by a more solid neighbour- is expressed in terms of φ while the flow can be considered
hood. Experimentally, the effect is particularly significant as almost incompressible (i.e. in the Boussinesq approxiclose to the jamming transition, when f vanishes. There- mation) if the non-local constitutive relation is expressed
fore, one needs to define a relative fluidity, which compares in terms of I. In the later case, the pressure P is instanthe degree of fluidity at one point and in its vicinity. As- taneously determined and becomes a state variable.
To conclude this section, we rewrite the non-local rhesuming as before, that the influence of fluidity is statistically isotropic and results from a short range interactions ology (12), with the choice of f = I = T γ̇ as a fluidity,
between shear zones, the relative fluidity can be defined as:
IY
as:
− I + ℓ2 ∇2 I.
(13)
0=
ℓ 2 ∇2 f
µ(I)
κ=
,
(11)
f
This formulation, directly derived from Eq.(12), makes it
where ℓ is a length on the order of few grain diameters. The easier to compare with the other approaches, as discussed
Laplacian is indeed the lowest order operator appearing in below.
6
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
3 The fluidity concept in soft elastic materials The standard Johnson-Segalman model corresponds to
Since fluidity was introduced to describe granular flows,
it is important to review the physical basis of the recent
advances made to render the complex rheology of soft matter (foams, gels, emulsions etc...) using this concept. Many
amorphous materials, including granular matter share the
same phenomenology, e.g. yield stress, Herschel-Bulkley
rheology. Although fluidity refers in standard rheology to
the inverse of viscosity, this name has been associated with
different quantities in the literature. In the recent conceptual picture derived from soft matter, fluidity appears as a
variable entering into the rheological constitutive relation
of the material. Qualitatively, low fluidity means closer
to a solid and large fluidity means closer to a fluid. Importantly, in those approaches fluidity dynamics obeys an
auxiliary equation which sets-in its temporal evolution.
This auxiliary equation reflects the microscopic or mesoscopic processes at the heart of the physics that one seeks
to describe. We shortly review the propositions made in
the literature to define fluidity.
3.1 Fluidity and elastic deformation
Many complex fluids are visco-elasto-plastic. It is the case
of polymer melts, micelles and lamellar surfactant phases,
which are subject to shear-banding, but also the case of
foams and emulsions, which are yield stress fluids. In this
situation, an obvious state variable controlling the rheology is the elastic strain ǫ. The elastic deformation is
an intrinsic property describing for example, the polymer
chains extension or for a foam, the current bubble deformation state. Importantly, ǫ is a field coarse grained in
space, in time, or averaged over realisations. The total
strain rate γ̇, determined from the subsequent positions
of the elements, can be formally written as the sum of the
elastic strain rate ǫ̇ and of a plastic contribution, equal
to γ̇ − ǫ̇. For convenience, we keep a scalar description
– the extension to a tensorial form does not present any
conceptual difficulty.
Non-locality in visco-elasto-plastic models was introduced by Olmsted [14] who has added non-local terms
terms to the Johnson-Segalman model [58]. As in the phase
field approach, the overall stress is decomposed into a
fluid-borne stress, for instance Newtonian η γ̇, and an elasticborne stress Gǫ, where η is a viscosity and G an elastic
shear modulus. The equation governing the evolution of
the deformation ǫ reads:
T ǫ̇ = T γ̇ − I(ǫ) + ℓ2 ∇2 ǫ,
(14)
where T ∼ η/G is the relaxation time-scale of the components (e.g. the polymer). The diffusion term ℓ2 ∇2 ǫ is
responsible for non-locality. The function I reflects the
elastic properties of the components and determines the
constitutive relation measured in a homogeneous steady
state. For the Herschel-Bulkley relationship (4) (we recall
that I = T γ̇), one obtains:
(1/n)
1
Gǫ
I(ǫ) =
.
(15)
−1
max 0,
a
σy
the exponent n = 1. The ratio ǫy = σy /G is the yield
strain above which plastic events nucleate. The same model
has successfully been used to describe complex fluids subject to shear-banding, with a vanishing yield stress [14,59].
In the steady state, the elastic deformation obeys the
non-linear Laplace equation ℓ2 ∇2 ǫ − I(ǫ) = −T γ̇, which
leads to exponential relaxations in space (see section 4).
In the limit where T becomes much smaller than γ̇ −1 ,
the same equation holds at all times and the dynamics
becomes truly non-local in the sense (i).
The above relation was modified by Marmottant and
Graner to model dry foams [60], assuming that the elastic
part does not evolve over to the internal time T but over
the time-scale γ̇ −1 , which is assumed to provide the only
time-scale of the problem. This assumption is close to that
made by Forterre and Pouliquen [56] for granular flows.
Keeping the structure of Eq. (14), the governing equation
then takes the form:
ǫ̇ = γ̇ 1 − I(ǫ) + ℓ2 ∇2 ǫ ,
(16)
where I is a function which is essentially 0 at low ǫ and
which sharply increases and crosses 1 at the yield strain
σy /G. Note that the derivation and tests proposed in [60]
actually do not take the non-local term into account (they
have ℓ = 0). We introduce it here, to clarify the connections between the various models.
3.2 Fluidity as a Maxwell relaxation rate
Considering again the ideal linear Couette cell controlled
at imposed shear stress σ, the mechanical response of a
soft material usually presents a transient in time. Over
time-scales comparable to γ̇ −1 or smaller, the structure
does not have time to evolve: the time dependence is said
to reflect visco-elasticity. If the rheology evolves over timescales long compared to γ̇ −1 , the system is said thixotropic.
Thixotropy is very close for time dependence to non-locality
for space dependence. It reflects ageing, a property often shared by many soft glassy materials arrested in the
glassy regime. Starting from a visco-elastic rheology relation, Derec et al. have proposed a model suited to describe such a thixotropic behaviour [51]. The fluidity f is
then introduced as a relaxation rate in a standard Maxwell
visco-elastic model :
ǫ̇ = −f ǫ + γ̇.
(17)
The auxiliary equation that governs the evolution of f is
macroscopic and implies ageing and shear rejuvenation
processes. By contrast, Olmsted model assumes a constant time for the relaxation and Marmottant and Graner
a time-scale inversely proportional to γ̇. The three models
thus differ by the identification of the relevant strain relaxation time. Interestingly, in the solid phase (i.e. below
the Coulomb threshold) sheared granular packing display
ageing that can be described by this equation and the
fluidity parameter f is directly related to the rate of localised plastic events, called hot-spots that were directly
visualised [61]. We will turn back to this point later.
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
7
Fig. 2. Evidences of Non-local behaviour for granular matter. (a) Experimental measurements of the stopping (black
symbols) and starting (white symbols) heights on an inclined plane with glass beads, data from [30]. The dependence
of stoppage and onset of flow with the granular height are manifestations of the non-local character of the rheology.
(b) Normalised creep velocity of an intruder in a Couette cell filled with glass beads as a function of the normalised
pulling force on the intruder, data from [47]. Data are taken in conditions where the intruder should be blocked (in
the solid phase) in a local rheology picture based on a Coulomb yield criterion. (c) Experimental velocity profile of a
dry granular avalanche flow in a narrow channel with frictional lateral walls. The dashed line is the prediction of the
local rheology displaying a depth z where flow stops whereas the granular continues to flow below this limit.
3.3 Fluidity as the rate of plastic events
Close enough to the yield stress, flow occurs in concentrated emulsions and foams via a succession of reversible
elastic deformations [62,63] and avalanches of irreversible
plastic rearrangements (also called shear transformation
zones in the literature [64]). Such localised plastic events
induce a long range anisotropic relaxation of the elastic
stress over the system, which constitutes an obvious source
of non-locality. Based on this observation, the fluidity in
a class of models has been associated to the rate of plastic
events, coarse grained in space and time. The physical picture assumes that the material is essentially in an elastic
state under stress but due to disorder or temperature or
mechanical fluctuations, localised plastic events nucleate,
which induce local irreversible stress relaxation processes.
The resulting deformation is absorbed elastically by the
medium. This is at the origin of a global plastic deformation rate. For a given fast and local relaxation process,
other plastic events may be triggered. When there is a
continuous rate of coupled plastic events spanning a significant time, this is called an avalanche [25,26,27,73]. This
avalanching process occurs preferentially at larger stress.
The exact account for avalanche dynamics as a net contribution to the final plastic deformation rate is a difficult
issue and has therefore been, in most models, phenomenologically modelled by means of non-linear terms entering
the fluidity equation.
The auxiliary equation for f can be a stochastic Smoluchowsky equation [65] but macroscopic – essentially meanfield – models have also been derived explicitly. For example Bocquet et al. [6] have proposed a model, called KEP
(kinetic elasto-plastic), which adds a spatial coupling to
the probabilistic framework introduced by Hébraud and
Lequeux [65]. The model describes the evolution of the
probability to observe locally a certain local ”shear stress”
and furthermore assumes that plastic events are triggered
above a non-fluctuating ”local yield stress”. The macroscopic rheology turns out to be controlled by the behaviour
of the shear stress probability distribution function in the
immediate vicinity of the local yield stress: in a steady
homogeneous state, a Herschel-Bulkley rheology is recovered.
In the KEP model, plastic events lead to a noise around
them that helps to trigger other plastic events. Non-locality
therefore appears as a dependence of a ”local stress diffusion” on the rate of plastic events around the region considered. Both the elastic strain (and therefore the stress)
and the rate of plastic events stem from an integral over
the same probability distribution function. In the steady
state (but not during transients), they can be related to
each other. One recovers the relation between the rate f
of plastic events and elastic strain of Eq. (17), for ǫ̇ = 0:
f=
γ̇
.
ǫ
(18)
The non-local equation has been derived by Bocquet et al.
in the steady state (and only in this case) and takes the
form of a non-linear Poisson equation, analogous to that
of Olmsted’s model (14) in the steady state:
0 = γ̇ − T −1 I(f ) + ℓ2 ∇2 f.
(19)
We later refer to this equation as the KEP constitutive
relation or KEP model. However, a mathematical analogy
is not a physical equivalence: the fluidity defined as the
variable that appears in the non-local Laplacian term is,
in one case, the elastic deformation ǫ and in the other, its
relaxation rate f . Furthermore, to fully solve the problem
in association with a Laplacian term in the formulation,
8
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
it is necessary to provide boundary conditions for fluidity
[66]. The choice of the expression for the fluidity must then
be consistent with the physical boundary conditions. We
will turn later on the possible tests to determine which
fluidity is the relevant one.
Importantly, the macroscopic emergence of the auxiliary fluidity equation involves a Laplacian operator which
physically represents the spatial range of plastic relaxations. Mathematically, it is the source of non-locality in
the constitutive relation. A possible issue is that the stress
relaxation induced by plastic events can have in general
an anisotropic character, even in a statistical sense, which
is not reflected by the isotropic Laplacian term in (19). If,
in this triggering process, anisotropy is important, higher
order terms must be included in the spatial expansion of
the stress propagator in [6]. Also, close to the yield condition, long range avalanches may take place and spatial
coupling can span large distances that eventually diverge
at the yield point. In this limit, the Laplacian, which is essentially a mean field operator, is unlikely to capture nonlocality, as it is well known that Landau-like approaches
generically fail in the vicinity of a critical point.
3.4 Fluidity as the inverse of viscosity
Bocquet et al. [6] have made a further, apparently innocent, step: as the elastic stress σ is proportional to the
elastic strain ǫ, the structure of the equations does not
change if f is defined as
γ̇
(20)
σ
instead of Eq. (18). The fluidity f becomes the inverse of
the particle-borne viscosity. The equation governing f remains of the form (19). We will discuss below the problems
associated with the change of variable ǫ → σ. We emphasise again that the relation between the rate of plastic
events and the fluidity is not trivial and must be tested.
To handle granular matter, Kamrin et al. [55] have
proposed to rescale this inverse viscosity by the yield stress
σy . If σy is constant, this does not change the shape of the
equation governing the new fluidity (Eq. 19):
σy
γ̇.
(21)
f=
σ
The fluidity f is then homogeneous to a strain rate. Again,
the function I can be determined from the constitutive
relation measured in a homogeneous steady state. For
the Herschel-Bulkley relationship, one obtains the implicit
equation:
I(f )
= T f,
(22)
1 + aI n (f )
whose solution takes the form:
f=
I(f ) = T f + a(T f )1+n + O(f 1+2n ).
(23)
For the particular case of frictional granular material, for
which n = 1, one gets the analytical solution:
I(f ) =
Tf
.
1 − aT f
(24)
4 What are the differences between non-local
constitutive relations proposed for granular
matter?
4.1 Fluidity, from soft-matter to granular matter
As seen above, the fluidity in Kamrin et al.’s model for
granular matter [55], derived from the results of Bocquet
et al. [6], writes
f=
γ̇
µc P
σy
γ̇ =
=
γ̇.
σ
Y
σ
(25)
The auxiliary equation for this parameter is derived from
Eq. (19) by a linearisation that we discuss below in details,
as it is problematic.
The transposition from soft matter to rigid grains poses
a fundamental issue. The fluidity must obviously be a state
variable. By state variable, we mean that the fluidity f
must be a coarse-grained field (in space and time) which
can be determined from the state of the system. It can for
example depend on the strain, the strain rate, the volume
fraction, the mean number of contacts. For a granular material, it can also involve the fraction of sliding contacts
or the orientation of the contacts. For foams or emulsions,
it can reflect the elastic deformation of elementary cells.
However, it cannot depend explicitly on the stress tensor,
which is not a state variable itself. This directly results
from Newton second’s law, which tells that positions and
velocities of the particles determines the state of a mechanical system, from which forces are derived. Similarly,
the (non-local) constitutive relation must relate the stress
tensor to the state variables – and not the opposite. f
as defined in Eq. (25) was a state variable for an elastic
system, because it was fundamentally based on the elastic
deformation. It is no longer the case for a granular system
composed of rigid grains.
4.2 Does flowing granular matter exhibit plastic
events?
The transposition of models derived for elasto-plastic material to granular systems obviously requires that granular
matter behaves as hypothesised by elasto-plasticity. The
elasto-plastic picture assumes that the material behaves
most of the time like a solid, but presents local and shortlived plastic events [64,26]. The associated scenario is a localised rupture initiation followed by a scale free avalanche
of localised events. In order to investigate whether this picture constitutes an alternative to the jamming scenario to
interpret the non-local nature of the granular rheology, we
compare, by means of numerical simulations (Discrete Element Method), the dynamics of two otherwise identical
systems composed of hard and soft grains.
The general numerical set-up is that used in [37]: we
consider a two-dimensional system composed of ∼ 2.103
spherical particles of a mean diameter d, with a ±20%
polydispersity. Such a choice ensures that the sample will
not crystalise, The particles can interact through contact
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
forces modelled as a viscoelastic force along the normal
contact direction and as a Coulomb friction along the tangential direction. The corresponding coefficient of restitution is e ≃ 0.9. The Coulomb friction coefficient is set to
µp = 0.4 for fictional particles and µp = 0 for the frictionless system. The particles are confined in a plane shear cell
composed of two rough solid walls made by the same particles, glued together. Periodic boundary conditions are
used along the shear direction x. The position of the wall
is controlled in order to impose a constant normal stress P
and constant and opposite velocities of the walls along x.
The system is in the asymptotic rigid limit when the ratio
kn /P of the normal spring constant with the pressure is
sufficiently large (typically above 103 ).
The presence of localised plastic events is usually based
on a visual inspection of different fields. The squared deviation from an affine deformation on a local scale has for
instance been proposed as a field indicating plastic activity [64,67]. However, such a quantity, as well as all those
based on the non-affine velocity, characterises fluctuations
around the mean flow, and not the local contribution of a
certain area to the mean flow. We wish here to propose a
practical definition of these events, based on the quantitative criterion that they must be separated in time and
localised in space. Importantly, to match their role played
in elasto-plastic models [64,67,6,68], they must also contribute additively to the average shear rate γ̇.
In order to detect localised plastic events, we have built
a coarse-grained field Γ̇ (r, t) reflecting, at time t, the local
contribution to γ̇ of a small region around the position r.
We impose that the time average of Γ̇ must everywhere
give γ̇. A coarse-graining method similar to that proposed
for the stress tensor [69,70] is adapted here to the computation of velocity differences and we take:
N
P
j=1
9
Fig. 3. (Color online) Image sequence showing the temporal evolution of the local shear rate Γ̇ for a system of
frictional hard grains (kn/P = 2 104 ) in the quasi-static
limit (I = 5 10−4 ). Similar results are found for frictionless particles. Color code from blue (Γ̇ = −40γ̇) to red
(Γ̇ = 40γ̇). Time lapse between two successive images:
γ̇∆t = 10−3 .
2
[ui (r, t) − uj (r, t)][zi (t) − zj (t)] exp( −||∆r||
)
2δ 2
(26)
, Gaussian distribution of hΓ̇ i around γ̇, while the PDF corresponding to the soft system shows stretched tails, which
[zi (t) −
are due to a very intermittent behaviour associated with
j=1
these plastic events. The PDF of δ Γ̇ provides informations
where ui (r, t) ispthe velocity of the grain i at the time about spatial heterogeneities in the system. The peak of
t and ||∆r|| = [zi (t) − zj (t)]2 + [xi (t) − xj (t)]2 denote the black line around 10γ̇ in Fig. 5d indicates that they
the distance between the grain i and j. δ is the coarse- are large and permanent in the hard system. For the soft
graining length, typically on the order of the grain size d. system, the PDF shows an algebraic decay, which means
We display in Fig. 3 and 4, for a system of rigid and soft that the field Γ̇ is homogeneous most of the time. However,
grains respectively, the map of the local contribution Γ̇ to when the computation of δ Γ̇ is restricted to the periods
the shear rate γ̇ at different times. In Fig. 5a,b we show of time where plastic events occur (periods where |hΓ̇ i| is
corresponding spatio-temporal diagrams built on the cen- larger than a given value, here 5γ̇ in Fig. 5d), its PDF also
tral line of the cell. We observe contrasted behaviours in presents a peak: in the soft system, plastic events are asthe two cases. In the soft system, nothing much happens sociated with a very heterogeneous field of Γ̇ . Conversely,
most of the time, except for short periods of intense activ- an assembly of rigid particles does not present local plasity, associated with a cascade of plastic events. Conversely, tic events when sheared permanently. Its dynamics is not
the hard system presents more moderate but permanent intermittent but presents spatial heterogeneities.
fluctuations even for asymptotically small γ̇.
In conclusion, when constituted of rigid particles, sheared
To make these observations quantitative, Fig. 5c shows granular systems do not present a succession of elastic
the probability distribution function (PDF) over time of energy accumulation and sudden release. Their dynamics
hΓ̇ i, which is the space average of Γ̇ over the cell. In panel rather show permanent cooperative motions. As a con(d), we similarly display the PDF of the spatial standard sequence, approaches explicitly based on elasto-plasticity
deviation δ Γ̇ . The hard-particle system presents a narrow developed for soft systems, such as those discussed above,
Γ̇ (r, t) =
N
P
2
)
zj (t)]2 exp( −||∆r||
2δ 2
10
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
0.10
0.10
0.05
0.05
0
0
0
7
14
21
28
0
35
100
7
14
21
28
35
100
10-1
10-1
10-2
10-3
10-2
10-4
10-3
-20
Fig. 4. (Color online) Image sequence showing the temporal evolution of the local shear rate Γ̇ for a system of
frictional soft grains (kn/P = 10) in the quasi-static limit
(I = 5 10−4 ). Similar results are found for frictionless particles. Color code from blue (Γ̇ = −40γ̇) to red (Γ̇ = 40γ̇).
Time lapse between two successive images: γ̇∆t = 10−3 .
cannot be transposed to granular flows, where elasticity of
the grains is irrelevant. The physical foundation Kamrin
et al.’s approach for granular matter [55] seems already in
this perspective, extremely problematic.
4.3 Linearisation in the case of a homogeneous shear
stress profile
The simplest situation in which the model predictions can
be tested is a shear cell inside which the shear stress,
and therefore the yield parameter Y, are homogeneous.
In such a situation, a local rheology predicts a constant
shear rate, which, once rescaled by T , is denoted I∞ . This
cell is driven by boundary layers on each side, whose properties are not necessarily those of the bulk. How to realise
this in practice for numerical simulations is for example
described in Bouzid et al. [37].
We first consider the case where I∞ does not vanish.
By definition, we have Y = µ(I∞ ). Making profit that Y
is constant, one can linearise the equations in I around
I∞ . Using the gradient expansion model (Eq. 13), where
I is the fluidity, we obtain:
n −1
ℓ2 ∇2 (I − I∞ ) = n (1 + aI∞
) − 1 (I − I∞ )
0
20
40
1
10
100
Fig. 5. (Color online) Space-time diagrams showing the
local contribution of Γ̇ to the shear rate γ̇, measured on
the central line of the cell, for a system of sheared hard
(a) and soft (b) grains. For this example, the ratio of the
grain contact stiffness kn to the overall pressure P is 2 104
and 10 respectively; the shear rate corresponds to an inertial number I = 5 10−4 . Color code from blue (Γ̇ = −40γ̇)
to red (Γ̇ = 40γ̇). (c) Probability distribution function
(PDF) over time of the space average hΓ̇ i, for hard (black
line) and soft (red line) grains. (d) PDF of the spatial standard deviation δ Γ̇ . Red dotted line: δ Γ̇ computed when
|hΓ̇ i| ≥ 5γ̇ (soft system).
=n
Y −1
(I − I∞ ).
Y
(27)
Denoting by z the axis transverse to the flow, we obtain
exponential solutions of the form
I = I∞ + A+ exp(z/L) + A− exp(−z/L),
(28)
where the relaxation length L is given by:
L2 =
n
ℓ2 Y
1 + anI∞
2
ℓ
=
n
anI∞
n(Y − 1)
for
Y > 1.
(29)
It is important to emphasise the status of this length L.
As Y is the control parameter of this particular thought
experiment (or numerical simulation [37]), L can be expressed as a function of Y. It does not make Y a state variable which would control another state variable L. Note
also that L can equally be expressed as a function of I∞ .
Consider now the KEP equation (19) with the fluidity
f proposed by Kamrin et al. (25). As above, making profit
that Y is homogeneous, this equation can be linearised
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
11
20
around f∞ as:
ℓ2 ∇2 (f − f∞ ) =
ℓ2
(f − f∞ ),
L2
where L is now given by:
1
1
2
2
+
L =ℓ
Y
n(Y − 1)
(30)
1
for
Y > 1.
(31)
0.9
In the limit Y → 1,pthe relaxation length takes exactly the
same form L ∼ ℓ/ n(Y − 1) for the two models, despite
their differences.
The second case corresponds to Y < 1, so that I∞ = 0.
Again, the equations can be expanded around I = I∞ but
this linearization is completely different from the previous
one, as I = 0 is not solution of the equations: one needs
the non-local term to get a solution. With the KEP model
(19), one obtains:
ℓ2 ∇2 f = (1 − Y)f,
(32)
which gives exponential relaxations over a length L given
by
ℓ2
for Y < 1.
(33)
L2 =
(1 − Y)
Using the gradient expansion model (13), we get:
ℓ2 ∇2 I = χ−1 (Y)I.
(34)
The shear rate therefore relaxes over a length L given by:
L2 =
ℓ2
χ−1 (Y)
for
1
Y < 1.
(35)
In the vicinity of the critical conditions, κ is indeed small,
so that the linear approximation χ(κ) ≃ 1−κ can be used.
The divergence of L at Y → 1 is therefore given, again,
by eq. (33).
As an illustration, we display in Fig. 6 such diverging
relaxation lengths extracted from numerical simulations
of sheared layer [37]. Velocity data, such a presented in
Fig.1b can be obtained systematically for Y values above
and below Y = 1. The fit of the velocity profiles is made
with a function of the form γ̇b z + C sinh(z/L), where C
and L are adjustable. This provides a direct measurement
of a relaxation length L, which effectively diverges on both
sides of the critical point Y = 1 according to the theoretical predictions (29) and (35). Fig. 7 shows the shape of
the function χ(κ). Its non-linear behaviour, which roughly
starts when κ > 0.1, is at the origin of the asymmetry of
L with respect to the yield point |Y| = 1 when sufficiently
far away from this point.
In conclusion, as far as linearisation is concerned, the
two models lead, despite their distinct starting point, exactly to the same predictions for constant stress conditions in the vicinity of the yield condition Y = 1. Because
two equations giving the same exponential solutions are
not necessarily equivalent, this shear cell configuration in
0.8
0.7
0
0.1
0.01
0.1
0.2
0.1
0.3
1
10
Fig. 7. Function χ(κ) numerically measured for Y < 1
for frictional (red circles) and frictionless (yellow circles)
grains – the system is that described [37]. The black
solid
line is a fit with the empirical expression χ =
√
(1−κα)2 +κβ(κα−1)
with α = −15.95 and β = 16.30. Inset:
1−ακ
zoom on the small values of κ, in Lin-Lin axes.
which the yield parameter Y is controlled and homogeneous thus cannot be used to discriminate between the different possibilities to build a fluidity. Further tests focusing on time transients and on heterogeneous situations are
needed to test the starting constitutive equations, which
do not reduce to their linearised expressions.
4.4 Does the KEP rheology reflect a dynamical phase
transition?
An important claim made by Bocquet et al. in [6] is that
the KEP constitutive relation reflects a dynamic phase
transition controlled by the stress, in relation to the divergence of the relaxation length L on both sides of Y = 1.
Because the model used by Kamrin et al. [55] is derived
from the KEP approach, this claim would also apply to
granular matter. On the opposite, in our framework, we
argue that the non-local rheology describes the same liquid phase above and below the yield conditions. Because
this controversy concerns an essential point of physics, we
find it important to develop in this subsection some technical but essential details of this issue.
As already mentioned, the starting KEP equation (19),
when linearised around a homogeneous stress state corresponding to a given constant Y, leads to the generic equation:
f − f∞
.
(36)
∇2 f =
L2
f∞ and L depend on the value of Y. This is the GinsburgLandau equation used by Kamrin et al. [55] with the fluidity given by Eq. (25). A crucial point is that this equation
is used by these authors in non-homogeneous situations,
i.e. taking for Y the local value, with functions f∞ (Y)
12
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
20
15
10
500.5
123.5
20
30
20
15
30
2
1
10
10
5
10
2
30
1
20
3
3
0.01
0.1
1
0.01
0.1
1
10
5
0
0
0
0.5
1
1.5
2
2.5
3
0
0.5
1
1.5
Fig. 6. Relaxation length L for frictionless (a) and frictional grains (c) below (red and yellow circles) and above (blue
and green squares) yield conditions, data from [37]. Solid lines: fit of the data to Eqs. (29) and (35), diverging as
|Y − 1|−1/2 when Y → 1. (b) and (d) Log-log plot of the same quantities.
and L(Y) [68,71]. This would be a correct assumption for
a slowly varying stress field if Y was a state variable. We
have discussed above why it is not the case, even though
it is the control parameter of the considered linearisation.
Let us further illustrate the mathematical differences
between two seemingly equivalent derivations. Consider
the KEP constitutive relation (19) with γ̇ = f Y (Eq. 25).
Associated with (22), which determines I(f ) in the homogeneous case, and whose solution is given by (24) for
n = 1, we obtain:
1
2 2
−Y .
(37)
ℓ ∇ f =f
1 − aT f
However, let us alternatively start from (36) and plug in
expressions of f∞ and L. f∞ is the solution of the equation γ̇ = T −1 I(f∞ ), i.e. f∞ (Y) = (Y − 1)/(aT Y). The
expression for L depends on whether Y is larger (Eq. 31)
or smaller (Eq. 33) than unity. To make it compact, let
us focus close to the yield condition Y = 1 for which both
cases can be summed up with L(Y)2 ∼ ℓ2 /|Y − 1|. Doing
so, instead of (37), we obtain:
Y −1
ℓ2 ∇2 f = |Y − 1| f −
.
(38)
aT Y
Equations (37) and (38) are obviously not the same. In
particular, their behaviour is very different when Y → 1
as the right hand side vanishes in the second case but
stays finite in the first one. In other words, the original
KEP equation (19) and the final Ginsburg-Landau equation used by Kamrin et al., are irreducible one to the
other: the transformation is neither mathematically nor
physically justified.
The divergence of the relaxation length L on both sides
Y = 1 has been interpreted as the signature of a dynamic phase transition controlled by Y. The region Y < 1
would correspond to the solid-like behaviour while the region Y > 1 would correspond to the liquid-like behaviour.
However, the original KEP model (37) was entirely based
on the description of a single liquid-like phase and the
Ginsburg-Landau equation (38) used by Kamrin et al. [55]
is not a controlled approximation of it. Y, which is not a
state variable, cannot be the parameter controlling the
phase transition. The liquid flowing phase can exist even
below the yielding conditions, for Y < 1.
4.5 Fluidity and boundary conditions
An important consequence of the choice of a particular fluidity f is the underlying assumption that f and its gradient ∇f are continuous, otherwise the use of the Laplacian
operator would not make any sense. This remark provides
a constraint on the nature of f , which can for example
be tested in a situation where the stress varies extremely
rapidly in space between two states. In this spirit, we
have performed numerical simulations where a secondary
micro-rheometer is placed in the bulk of a shear cell as presented previously (see Figs.1b and 2 and details in Bouzid
et al. [74]). Shearing within the micro-rheometer is obtained by means of localised bulk forces along two lines
which induce a discontinuity of the shear stress. We have
measured numerically the ratio R of the absolute value of
the shear rate on one side and on the other side of the
stress discontinuity. The pressure remains constant and
the direction of shearing is reversed ( γ̇ changes sign at the
discontinuity). Fig. 8 shows that |γ̇| is indeed continuous
(R = 1) in both frictionless and frictional cases. On the
opposite, the fluidity proposed by Kamrin et al. (Eq. 25) is
not in agreement with the data, especially when the yield
parameter approaches zero.
5 Further tests: the question of dynamical
mechanisms
We have presented a short review of non-locality in granular flows. We have mainly focussed on the comparison of
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
1
0.1
-1
-0.8
-0.6
-0.4
-0.2
0
Fig. 8. Ratio R of the absolute value of the shear rate
on the outer and inner sides of a stress discontinuity, as
a function of the yield parameter Ym . Yellow circles are
for frictionless grains and gray lozenges are for frictional
grains. Thick solid line R = 1: prediction of the gradient
expansion model [37] using I as fluidity. Dotted line: prediction of fluidity theory [55] using f = γ̇/σ as a fluidity
parameter.
two models: the KEP model adapted for granular matter
by Kamrin et al.[55,68] (see Eq. 25), and the gradient expansion that we have proposed [37] (see Eqs. 12 and 13)).
The difference between these models has not been recognised so far in the literature, essentially because they lead
to the same predictions for the velocity profile in a situation where the stress is homogeneous. However, our conclusion is that these models are fundamentally different
and that their differences can be tested.
These tests must be performed in situations that are
strongly heterogeneous in space, like the one shown in
Fig. 8, or to unsteady situations [75]. They can of course
concern the direct predictions of the model, but also the
choice for the fluidity, which is not necessarily the inverse
of viscosity, as well as the associated boundary conditions.
Importantly, these conditions should not be fitted, but
part of the physical analysis of the problem. More fundamentally, the difference between the models is to be found
in the hypothesis made to derive them and in particular
in the dynamical mechanisms underlying their dynamics.
For instance, many different explanations have been proposed for the very same Herschel-Bulkley rheology. In the
context of non-locality, let us give several example where
ingredients could be tested. Different models assume the
proportionality between the decay rate of fluidity and the
rate of plastic events. Such a proportionality can therefore
be investigated experimentally. Other models like KEP
prescribe not only the average fluidity but also its distribution. The measurement of such a quantity is a more
severe test than the fit of velocity profiles. Would a model
assume the existence of microscopic yield conditions for
the nucleation of plastic events, it would then be necessary to determine this quantity, to show that it exists and
that it is constant.
13
In the case of granular matter, we have shown that the
main point separating the KEP model and the gradient
expansion model, is the existence or not of elasto-plastic
localised events in the liquid regime. The KEP model is
directly adapted from soft matter and assumes that elasticity dominates the dynamics. In the test presented here,
we have numerically shown that, in the rigid limit, there
are no localised plastic events and the flow is dominated by
non-affine collective motion along soft modes. One could
argue that the Coulomb friction condition at the contacts
between the grains may lead to plastic events. However,
comparing frictional and frictionless grains, we do not see
any difference neither on non-locality nor on the absence
of plastic events. Beyond other important reasons, we have
shown that the fluidity proposed by Kamrin et al. [55], as
an extension of the KEP model to rigid granular packings,
is not a state variable and is thus not continuous across a
stress discontinuity. We acknowledge that, in spite of the
fact that our choice of the fluidity parameter for dense
granular flows respects the state variable requirements and
quantitatively predicts some situations, it does not clarify
the understanding of the actual microscopic mechanisms
at work to definitely unravel the question of non-local rheology. We recently discussed some limitations on such a
choice and we pursue the work of identifying the microscopic or mesoscopic processes associated with the flow of
hard grains [74,75].
Finally, amongst the points that have created a confusion in the literature, is the fact that granular matter
does present localised plastic events, but only in the solid
regime [61,76], not in the liquid regime discussed here,
were grain elasticity is irrelevant. Let us note that numerical simulations performed with the standard Coulomb
model of friction at contacts, which perfectly reproduce
observations in the dense liquid regime and in particular
non-locality, are not able to reproduce creep in the solid
regime. These two regimes (solid-like and liquid-like) must
eventually be described, but the transition between these
dynamical phases is known to be subcritical and to present
a hysteresis, a key aspect of granular matter that remains
unexplained at present. In this context, a non-local transition between solid and liquid was addressed by Wyart
[77], based on the generic outcome of the Maxwell rigidity transition for hard granular packing. Importantly, the
KEP model claims to describe this dynamical phase transition as a critical transition controlled by the stress, and
the rheology both above and below the transition. The
gradient expansion, on the opposite, is based on the fact
that shear stress cannot be a control parameter for this
transition and describes the system as a unique continuous
liquid phase both above and below yield conditions. This
approach is therefore perfectly compatible with a subcritical transition to the solid regime, as it does not describe
the later.
References
1. B. Andreotti and S. Douady, Phys. Rev. E 63, 031305
(2001).
14
2.
3.
4.
5.
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
J. Rajchenbach, Phys. Rev. Lett. 90, 144302 (2003).
J. T. Jenkins, Phys. Fluids 18, 103307 (2006).
B. Andreotti, Europhys. Lett. 79, 34001 (2007).
J. Goyon, A. Colin, G. Ovarlez, A. Ajdari and L. Bocquet,
Nature. 454, 84 (2008).
6. L. Bocquet, A. Colin and A. Ajdari, Phys. Rev. Lett. 103,
036001 (2009).
7. J. Goyon, A. Colin and L. Bocquet, Soft Matter 6, 2668
(2010).
8. P. Chaudhuri, V. Mansard, A. Colin and L. Bocquet, Phys.
Rev. Lett. 109, 036001 (2012).
9. V. Mansard and A. Colin, Soft Matter 8, 4025 (2012).
10. I.S. Aranson and L.S. Tsimring, Phys. Rev. E 64,
020301(R) (2001).
11. I.S. Aranson and L.S. Tsimring, Phys. Rev. E 65, 061303
(2002).
12. D. Volfson, L.S. Tsimring and I.S. Aranson, Phys. Rev. E
68, 021301 (2003).
13. I.S. Aranson, L.S. Tsimring, F. Malloggi and E. Clément,
Phys.Rev.E 78, 031303 (2008).
14. P. D. Olmsted, Rheol. Acta 47, 283 (2008).
15. M. Trulson, M. Bouzid, J. Kurchan, E. Clément, P.
Claudin, and B. Andreotti, Europhys. Lett. 111, 18001
(2015).
16. M.D. Ediger, C.A. Angell and S.R. Nagel, J. Phys. Chem.
100, 13200 (1996).
17. P. Debenedetti and F. Stillinger, Nature 410, 259 (2001).
18. R. Mari, F. Krzakala and J. Kurchan, Phys. Rev. Lett.
103, 025701 (2009).
19. P. Olsson and S. Teitel, Phys. Rev. Lett. 99, 178001(2007).
20. P.N. Pusey and W. van Megen, Nature 320, 340 (1985).
21. T.G. Mason and D.A. Weitz, Phys. Rev. Lett. 75, 2770
(1995).
22. G.L. Hunter and E.R. Weeks, Rep. Prog. Phys. 75, 066501
(2012).
23. J. Mewis and N. Wagner, Colloidal Suspension Rheology,
Cambridge University Press, New York (2012).
24. M. Cloı̂tre, R. Borrega, F. Monit and L. Leibler, Phys.
Rev. Lett. 90, 068303 (2003) .
25. D.S. Fisher, Physics Reports 301, 113 (1998).
26. A. Tanguy, F. Leonforte and J.-L. Barrat, Eur. Phys. J E
20, 355 (2006).
27. H.G.E. Hentschel, S. Karmakar, E. Lerner and I. Procaccia, Phys. Rev. Lett. 104, 025501 (2010).
28. B. Andreotti, Y. Forterre and O. Pouliquen, Granular Media: Between Fluid and Solid, Cambridge University Press
(2013).
29. A.J. Liu and S.R. Nagel, Nature 396, 21 (1998).
30. GDR MiDI, Eur. Phys. J. E. 14, 341 (2004).
31. F. da Cruz, S. Emam, M. Prochnow, J.N. Roux and F.
Chevoir, Phys. Rev. E 72, 021309 (2005).
32. M. Trulsson, M. Bouzid, P. Claudin and B. Andreotti Europhys. Lett. 103, 38002 (2013).
33. O. Pouliquen, Phys. Fluids. 11, 542 (1999).
34. P. Jop, Y. Forterre and O. Pouliquen, Nature. 441, 727
(2006).
35. F. Boyer, E. Guazzelli and O. Pouliquen, Phys. Rev. Lett.
107, 188301 (2011).
36. M. Trulsson, B. Andreotti and P. Claudin, Phys. Rev. Lett.
109, 118305 (2012).
37. M. Bouzid, M. Trulsson, P. Claudin, E. Clément and B.
Andreotti, Phys. Rev. Lett. 111, 238301 (2013).
38. G. Katgert, B.P. Tighe, M.E. Möbius and M. van Hecke,
Europhys. Lett. 90, 54002 (2010).
39. J.T. Jenkins and S.B. Savage, J. Fluid Mech. 130, 187
(1983).
40. T.S. Komatsu, S. Inagaki, N. Nakagawa and S. Nasuno,
Phys. Rev. Lett. 86, 1757 (2001).
41. N. Taberlet, P. Richard, A. Valance, W. Losert, J.M.
Pasini, J.T. Jenkins and R. Delannay, Phys. Rev. Lett. 91,
264301 (2003)
42. N. Taberlet, P. Richard and R. Delannay, Computers and
Mathematics with Applications 55, 230 (2008).
43. V.B. Nguyen, T. Darnige, A. Bruand and E. Clément,
Phys. Rev. Lett. 107, 138303 (2011).
44. F. Malloggi, B. Andreotti and E. Clément, Phys. Rev. E
91, 052202 (2015).
45. S. Deboeuf, E. Lajeunesse, O. Dauchot and B. Andreotti,
Phys. Rev. Lett. 97, 158303 (2006).
46. K. Nichol, A. Zanin, R. Bastien, E. Wandersman and M.
van Hecke, Phys. Rev. Lett. 104, 078302 (2010).
47. K.A. Reddy, Y. Forterre and O. Pouliquen, Phys. Rev.
Lett. 106, 108301 (2011).
48. E. Wandersman, M. Van Hecke, Europhys. Lett. 105,
24002 (2014).
49. R. Harich, T. Darnige, E. Kolb and E. Clément, Europhys.
Lett 96, 54003 (2011).
50. H.B. Mühlhaus and I. Vardoulakis, Géotechnique 37, 271
(1987)
51. C. Derec, A. Ajdari and F. Lequeux, Eur. Phys. J. E 4,
355 (2001).
52. F. Malloggi, J. Lanuza, B. Andreotti and E. Clément, Europhys. Lett. 75, 825 (2006).
53. I.S. Aranson, F. Malloggi, and E. Clément, Phys. Rev. E
73, 050302(R) (2006).
54. E. Clément, F. Malloggi, B. Andreotti, I.S. Aranson, Granular Matter 10, 3 (2007).
55. K. Kamrin and G. Koval, Phys. Rev. Lett. 108, 178301
(2012).
56. O. Pouliquen and Y. Forterre, Phil. Trans. R. Soc. A. 367,
5091 (2009).
57. M. Bouzid, PhD Thesis, Comportement rhéologiques et effets non-locaux dans les écoulements granulaires denses. Université Paris Diderot (2014).
58. M. Johnson and D. Segalman, J. Non-Newt. Fluid Mech.
2, 255 (1977).
59. P. D. Olmsted, O. Radulescu and C.-Y. D. Lu Jou. Rheol.
44, 257 (2000).
60. P. Marmottant, F. Graner Eur. Phys. J. E. 23, 337 (2007).
61. A. Amon, V. B. Nguyen, A. Bruand, J. Crassous, and E.
Clément, Phys. Rev. Lett. 108, 135502 (2012).
62. D. Chen, K.W. Desmond and E.R. Weeks, Phys. Rev. E
91, 062306 (2015).
63. K.W. Desmond and E.R. Weeks, Phys. Rev. Lett. 115,
098302 (2015).
64. M.L. Falk and J.S. Langer Phys. Rev. E 57, 7192 (1998).
65. P. Hébraud and F. Lequeux, Phys. Rev. Lett. 81, 2934
(1998).
66. V. Mansard, L. Bocquet and A. Colin, Soft Matter 10,
6984 (2014).
67. A. Lemaı̂tre and C. Caroli, Phys. Rev. Lett. 103, 065501
(2009).
68. D.L. Henann and K. Kamrin, Proc. Natl. Acad. Sci. USA
110, 6730 (2013).
Mehdi Bouzid et al.: Non-local rheology in dense granular flows
69. B.J. Glasser and I. Goldhirsch, Phys. Fluids 13, 407
(2001).
70. C. Goldenberg and I. Goldhirsch, Phys. Rev. Lett. 89,
084302 (2002).
71. D.L. Henann and K. Kamrin, Phys. Rev. Lett. 113, 178001
(2014).
72. R. Benzi, M. Sbragaglia, A. Scagliarini, P. Perlekar, M.
Bernaschi, S. Succi and F. Toschi, Soft Matter 11, 1271
(2015).
73. K.A. Dahmen, Y. Ben-Zion, and J.T. Uhl, Nature Phys.
7, 554 (2011).
74. M. Bouzid, M. Trulsson, P. Claudin, E. Clément and B.
Andreotti, Europhys. Lett. 109, 24002 (2015).
75. E. Rojas, M. Trulsson, B. Andreotti, E. Clément and R.
Soto, Europhys. Lett. 109, 64002 (2015).
76. A. Le Bouil, A. Amon, S. McNamara, and J. Crassous,
Phys. Rev. Lett. 112, 246001 (2014).
77. M. Wyart, Europhys. Lett. 85, 24003 (2009).
15