Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Interplay between hysteresis and nonlocality during onset and arrest of flow in granular materials

Saviz Mowlavi and Ken Kamrin *
Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge, MA 02139, USA. E-mail: kkamrin@mit.edu

Received 5th May 2021 , Accepted 2nd July 2021

First published on 8th July 2021


Abstract

The jamming transition in granular materials is well-known for exhibiting hysteresis, wherein the level of shear stress required to trigger flow is larger than that below which flow stops. Although such behavior is typically modeled as a simple non-monotonic flow rule, the rheology of granular materials is also nonlocal due to cooperativity at the grain scale, leading for instance to increased strengthening of the flow threshold as system size is reduced. We investigate how these two effects – hysteresis and nonlocality – couple with each other by incorporating non-monotonicity of the flow rule into the nonlocal granular fluidity (NGF) model, a nonlocal constitutive model for granular flows. By artificially tuning the strength of nonlocal diffusion, we demonstrate that both ingredients are key to explaining certain features of the hysteretic transition between flow and arrest. Finally, we assess the ability of the NGF model to quantitatively predict material behavior both around the transition and in the flowing regime, through stress-driven discrete element method (DEM) simulations of flow onset and arrest in various geometries. Along the way, we develop a new methodology to compare deterministic model predictions with the stochastic behavior exhibited by the DEM simulations around the jamming transition.


1 Introduction

Granular materials are well-known for displaying both solid-like and fluid-like behavior depending on their internal stress state.1–3 Flow can be induced or arrested through external loading variations, which has direct implications for a wide range of catastrophic geophysical phenomena such as landslides, avalanches and earthquakes.4–6 The transition between solid-like and liquid-like behavior in frictional granular media is characterized by several unique macroscopic features, which have been uncovered through simple experiments in model systems.7Fig. 1 showcases typical results from such experiments, where flow is triggered then arrested by ramping up and down the applied stress in (a) an annular shear cell,8 (b) a layer of grains on an inclined plane,9 and (c) a partially-filled rotating drum.10 The features revealed in these experiments are universal to most geometries and can be outlined as follows:
image file: d1sm00659b-f1.tif
Fig. 1 Previous experimental investigations of the flow threshold in various geometries. (a) Annular shear cell:8 ratio of shear stress to pressure at the inner wall, μw, versus dimensionless mean strain rate, image file: d1sm00659b-t1.tif, for increasing and decreasing torque applied to the inner cylinder. (b) Inclined plane:9 angles of inclination at flow onset and arrest, θstart and θstop, versus dimensionless layer thickness, H/d. (c) Rotating drum:10 angles of inclination at flow onset and arrest, θstart and θstop, versus dimensionless drum width, W/d.

(F1) the level of stress required to trigger flow is larger than that below which flow stops, leading to a hysteresis of the flow velocity as the applied stress is ramped up and down;

(F2) the onset of flow is accompanied by a finite jump in the velocity of the system;

(F3) the critical stresses for onset and arrest of flow depend on the size of the system, with smaller system sizes displaying increased strengthening.

Each feature is directly relevant to geophysical events such as landslides and avalanches, since (F1) controls the mobilized mass that flows down, (F2) explains why they are so spontaneous and catastrophic, and (F3) determines the circumstances under which they might occur. The objective of the present work is to formulate a continuum model that is able to describe quantitatively the onset and arrest of flow in frictional granular materials in various two-dimensional geometries, and analyze how its constituent ingredients play a role in reproducing each of these three features, with particular focus on geometries displaying inhomogeneous flow fields.

It is now well accepted that dense and homogeneous flows of grains follow the μ(I) constitutive relationship, which states that the stress ratio μ and the inertial number I are related through a one-to-one function μ = μloc(I).7,11 In two dimensions, μ = τ/P is the ratio of shear stress τ to pressure P, and image file: d1sm00659b-t2.tif is the strain rate [small gamma, Greek, dot above] nondimensionalized with a particle-wise rearrangement time scale formed by the mean grain mass m and confining pressure P. While μloc(I) has long been believed to be a monotonic function of I, several recent experiments12–15 and simulations16,17 have revealed the existence at very low I of a strain-rate weakening regime, wherein the stress ratio μ decreases with increasing I. One possible non-monotonic functional form is

 
image file: d1sm00659b-t3.tif(1)
where μs, μ2, and b are dimensionless rheological constants, κ = kn/P a dimensionless stiffness with kn the grain stiffness, and χ is a decreasing function of I that accounts for the strain-rate weakening regime. The microscopic origin of the strain-rate weakening regime has recently come under debate, with some studies arguing that it is caused by inertia of the grains,17–19 while others observing velocity-weakening behavior in over-damped, inertia-less particulate media.14 The present work is not concerned with this particular issue, and we simply leave open the possibility for the amount of strain-rate weakening to depend on the grain stiffness through the dimensionless parameter κ entering χ. In any case, the non-monotonicity of (1) necessarily implies that features (F1) and (F2) above are realized in homogeneous flows: the level of stress required to trigger flow is higher than that at which flow stops, and flow onset is characterized by a velocity jump.20,21

However, the μ(I) constitutive relationship breaks down in inhomogeneous flows, in the sense that μ is no longer a one-to-one function of I.22,23 Due to the finite size of the grains, velocity fluctuations generated at an arbitrary location will spread over some grain-size-dependent correlation length and change the rheology of the neighboring material,24–27 resulting in wider shear regions than are predicted by the μ(I) rheology, especially in the quasi-static limit.7 Such spatial cooperativity at the scale of individual grains also explains why thinner layers on an inclined plane start flowing at higher inclination angles than thicker layers, despite the stress ratio μ being independent of the layer height.7 Nonlocal rheological models, which incorporate an intrinsic length scale, have been shown to capture several of these phenomena.28 Here, we focus on the nonlocal granular fluidity (NGF) model,29,30 which relates the stress ratio and strain rate through a granular fluidity field g = [small gamma, Greek, dot above]/μ that is governed by a reaction–diffusion partial differential equation (PDE). Our choice of the NGF model stems from its ability to reproduce the system-size dependence of the flow threshold in various geometries,31,32 which partially explains feature (F3) above. But the current formulation of the NGF model reduces to the monotonic form of the μloc(I) relationship in homogeneous flow conditions, meaning that the model does not have a built-in mechanism to account for the remaining features (F1) and (F2).

In this paper, we modify the NGF model so that it instead reduces to the non-monotonic form of the μloc(I) relationship, eqn (1), in homogeneous flows. By computing time-dependent model predictions in a stress-driven planar shear configuration under gravity, we evaluate the specific ways in which nonlocality and non-monotonicity contribute to each of the three features (F1–F3) of the flow-arrest transition in inhomogeneous flows. We show that inclusion of, and interplay between both ingredients is necessary to reproduce all three features, in ways sometimes surprising: the planar shear with gravity configuration displays a finite velocity jump during onset of flow only when both non-monotonicity and nonlocality are present. In a second part, we assess the capability of the modified NGF model to predict quantitatively the behavior of dense granular materials both around the flow-arrest transition as well as in the flowing regime. To this effect, we calibrate the model using discrete element method (DEM) simulations in the simple shear geometry shown in Fig. 2(a), and we compare predictions of the calibrated model against stress-driven DEM simulations in the other geometries displayed in Fig. 2(b) and (c), namely plane shear under gravity and inclined plane.


image file: d1sm00659b-f2.tif
Fig. 2 Geometries considered in this study: (a) plane simple shear, (b) plane shear under gravity, and (c) inclined plane flows. Particles that are free to flow are colored according to their relative velocity magnitude, and fixed wall particles are shown in brown.

These two configurations are both subject to nonlocal effects as a result of the spatial inhomogeneity of their flow fields, but they critically differ in an important way – in plane shear with gravity, flow inhomogeneity is mostly a consequence of the spatial dependence of the stress ratio μ, while it is mainly caused by the rough base in inclined plane flow.32 We observe that the accuracy of the NGF model depends on which of these two mechanisms is at play, with predictions being accurate in the case of plane shear with gravity but less so for inclined plane, which we ultimately attribute to the role played by the boundary conditions.

The remainder of this paper is organized as follows. In Section 2, we present a modified NGF model that incorporates a non-monotonic local rheology, and we evaluate the combined effects of nonlocality and non-monotonicity on the features of the flow-arrest transition. We then compare in Section 3 predictions from the NGF model with DEM simulation results in stress-driven planar shear with gravity and inclined plane configurations. We close the paper with concluding remarks in Section 4.

2 Nonlocal granular rheology

In this section, we discuss our nonlocal continuum modelling approach based on the nonlocal granular fluidity (NGF) model. We begin by presenting the NGF model in its current form, which does not capture the hysteresis of the flow-arrest transition. We then describe the incorporation of bistable behavior into the NGF model. Finally, we evaluate the combined effects of bistability and nonlocality on the qualitative behavior of the flow-arrest transition.

2.1 Nonlocal model without hysteresis

Extending earlier fluidity-based nonlocal models for concentrated emulsions25,33,34 to granular materials, the NGF model introduces a positive granular fluidity field g that relates the strain rate [small gamma, Greek, dot above] with the stress ratio μ through the following two constitutive equations:
 
[small gamma, Greek, dot above] = ,(2a)
 
image file: d1sm00659b-t4.tif(2b)
where t0 is a constant timescale associated with the dynamics of g, and the nonlocal amplitude A > 0 is a dimensionless scalar parameter quantifying the strength of spatial cooperativity in the flow. As a side note, we mention that recent studies35–37 have endowed the granular fluidity field with a clear physical meaning – g is a purely kinematic quantity related to the velocity fluctuations δv, grain size d and solid fraction ϕ through g = (δv/d)F(ϕ), where F depends solely on ϕ.

The flow rule (2a) states that the strain rate [small gamma, Greek, dot above] is directly proportional to the fluidity g. Therefore, there can only be flow provided g is nonzero, and the nonlocal granular rheology is driven by the dynamics of eqn (2b) for the fluidity. The latter takes the form of a reaction–diffusion equation, and its behavior can be understood as follows.

In the absence of boundary effects or nonuniformities in the stress ratio μ, the g field becomes spatially uniform and (2b) reduces to the simple dynamical system

 
image file: d1sm00659b-t5.tif(3)
where F(g;μ,P) is a simple linear function of g. The steady behavior of the system is then governed by the steady-state solutions gloc of (3), which we illustrate by the thick lines in Fig. 3 using arbitrary values§ for μs = 0.25, μ2, b, and P. When μμs, the function F(g;μ,P) < 0 for all g ≥ 0 as shown in Fig. 3(a), hence (3) only admits the stable, arrested steady state gloc = 0, represented by the horizontal part of the green branch in Fig. 3(b). When μ > μs, the linear function F(g;μ,P) acquires one positive root and F(g = 0;μ,P) > 0, which has two consequences. First, the arrested steady state becomes unstable, shown by the red branch in Fig. 3(b). Second, a stable, flowing steady-state solution gloc(μ) > 0 emerges, represented by the positive green branch in Fig. 3(b). Using the flow rule (2a) together with the definition of the inertial number I, the flowing and arrested stable solutions can be inverted and expressed as
 
image file: d1sm00659b-t6.tif(4)
for I > 0, and μlocμs otherwise. This relationship is pictured in green in Fig. 3(c), together with the unstable arrested solution above μs in red. Therefore, the NGF model reduces to the local μ(I) rheology in steady and homogeneous flows such as plane shear without gravity.


image file: d1sm00659b-f3.tif
Fig. 3 Steady-state solutions of the local limit of the NGF model for homogeneous flows, without hysteresis (thick lines) and with hysteresis (thin lines). (a) Behavior of F(g;μ,P) and Fh(g;μ,P) in (4) and (9), showing the existence of zero, one or two roots for different stress ratios μ. (b) Unstable (red) and stable (green) steady-state local solutions gloc as a function of μ. (c) Resultant local rheology μloc(I), with both the unstable (red) and stable (green) branches shown. The model with hysteresis displays a strain-rate weakening regime absent in the model without hysteresis.

In the presence of boundary effects or nonuniformities in the stress ratio μ, however, the diffusion term in (2b) spreads the granular fluidity over a cooperativity length scale proportional to the grain size d, resulting in a nonlocal flow rule (2a). Regions where μ > μs act as stress-driven sources of granular fluidity. The fluidity is then diffused towards lower-stress regions or boundaries. Such nonlocal, cooperative effects have manifold consequences, and the NGF model explains many phenomena evading local rheological models. For instance, the model recovers the decaying motion of grains in regions where μ < μs23,30,38 as well as the so-called secondary rheology, wherein flow anywhere in a granular media removes the yield stress elsewhere.39,40 Conversely, the model is able to explain the strengthening of the flow threshold with decreasing system size,31,32 which is caused by boundaries or low-stress regions preventing flow in other regions where μ > μs unless μ is large enough. This last property relates to feature (F3) mentioned in the introduction. Yet, the current form of the NGF model is unable to reproduce features (F1) and (F2) due to the monotonicity of its limiting local rheology (4).

2.2 Nonlocal model with hysteresis

We now discuss the inclusion of non-monotonicity of the local rheological response into the NGF model. Taking inspiration from previous hysteretic nonlocal models,41,42 we add a new term to the right-hand side of the fluidity eqn (2b) giving the following updated set of fluidity equations:
 
[small gamma, Greek, dot above] = ,(5a)
 
image file: d1sm00659b-t7.tif(5b)
where the new term χ(g;μ,P) takes the form
 
image file: d1sm00659b-t8.tif(6)
with a, c, n constant scalar parameters, and κ = kn/P the nondimensional particle stiffness. Here, we choose to express the new term χ with a tanh function so that it vanishes for large g, so that its influence is restricted to zones near the jamming transition.

As before, we begin by evaluating the dynamics of the g field for the case of homogeneous flows, in which g is spatially uniform and (5b) reduces to the simple dynamical system

 
image file: d1sm00659b-t9.tif(7)
where Fh(g;μ,P) is a function of g. Contrary to the previous case without hysteresis, the presence of the χ term induces a decrease in Fh(g;μ,P) when g approaches zero, as illustrated by the thin lines in Fig. 3(a) using the same parameter values as before and new arbitrary values for the parameters of χ. As a result, there exists a range of stress ratios image file: d1sm00659b-t10.tif in which the function Fh(g;μ,P) inherits a second positive root, so that (7) admits two flowing steady-state solutions gloc(μ) > 0, one stable and one unstable, shown by the thin lines in Fig. 3(b). These two flowing solution branches merge at μ = μ*. The stable branch reverts for image file: d1sm00659b-t11.tif to the same flowing solution as the NGF model without hysteresis, while the unstable branch merges at image file: d1sm00659b-t12.tif with the arrested steady-state solution gloc = 0, which remains stable until μ exceeds image file: d1sm00659b-t13.tif. Collecting the pieces, the steady-state solutions of (7) can be expressed in terms of the inertial number I as
 
image file: d1sm00659b-t14.tif(8)
for I > 0, and image file: d1sm00659b-t15.tif otherwise. The function χ is now formulated in terms of I and κ as
 
χ(I;κ) = a[1 − tanh(cIκn)],(9)
and the static yield stress ratio image file: d1sm00659b-t16.tif is obtained as
 
image file: d1sm00659b-t17.tif(10)
Thus, the modified NGF model reduces to the non-monotonic μ(I) rheology (1) in steady homogeneous flows, and the corresponding stable and unstable branches are displayed by the thin green and red lines in Fig. 3(c). In the quasi-static, low-I regime, the presence of χ induces a weakening relationship between μ and I. For higher values of I, the vanishing of χ leads to a strain-rate strengthening regime in which the non-monotonic local rheology (8) converges to its monotonic counterpart (4). The crossover between the two regimes occurs at dμloc/dI = 0, which corresponds to
 
image file: d1sm00659b-t18.tif(11)
In agreement with force balance arguments,20 the strain-rate weakening regime is unstable while the strain-rate strengthening regime is stable. Since the latter exists for μ above μ* and the arrested solution is stable below image file: d1sm00659b-t19.tif, there exist two stable steady-state solutions – one flowing and one arrested – in the range image file: d1sm00659b-t20.tif. In the absence of flow gradients, this bistable behavior generates hysteresis when the stress ratio μ is ramped up and down: flow is triggered at image file: d1sm00659b-t21.tif but stops at a lower μ*. In addition, the onset of flow is accompanied by a finite jump in the velocity of the system, as the inertial number jumps from the arrested solution to the stable flowing solution. Hence features (F1) and (F2) are accounted for, but it is unclear whether this would hold for inhomogeneous flow, in the presence of nonlocal diffusion imparted by boundaries or nonuniformities in the stress ratio.

2.3 Interplay between hysteresis and nonlocality

We now investigate qualitatively the combined effects of non-monotonicity and nonlocal diffusion on the characteristics of the flow-arrest transition in the presence of a spatially-varying stress ratio. To do so, we calculate quasi-steady, stress-driven predictions of the NGF model with hysteresis in the plane shear under gravity configuration pictured in Fig. 2(b), where flow occurs along the x-direction and gravity acts orthogonally along the z-direction. A shear stress τw and pressure Pw are applied at the top wall, imparting under quasi-steady conditions a constant shear stress τ(z) = τw and a nonuniform pressure P(z) = Pw + ϕρsGz, where ρs is the grain density, ϕ the mean area packing fraction and G the acceleration of gravity. The ratio of shear stress to pressure is thus given by
 
image file: d1sm00659b-t22.tif(12)
where μw = τw/Pw is the applied stress ratio at the top wall, and [small script l] = Pw/ϕρsG is a loading length scale measuring the relative importance of the pressure imparted by the top wall versus that due to the weight of the grains. Critically, [small script l] is inversely proportional with the degree of nonuniformity of the stress ratio (12) and, thus, the strength of nonlocal effects in this geometry.43 Results from previous DEM simulations32 as well as our own (see Section 3.2) have shown that these nonlocal effects induce the same flow-arrest transition features (F1–F3) that are observed in other geometries.

We compute quasi-steady, time-dependent solutions of the NGF model with hysteresis using the same arbitrary parameters as in the previous section. Because the dynamics are uniform in the streamwise x-direction, the fluidity eqn (5b) reduces to a one-dimensional PDE for g(z,t), which is discretized following the procedure presented in Appendix B. From there, the strain rate [small gamma, Greek, dot above](z,t) and therefore the velocity profile v(z,t) can be computed using the flow rule (5a). The fluidity equation is driven by the stress ratio (12), for which we choose an arbitrary value [small script l] = 100d small enough that the results are independent of the height of the domain. Following previous work,32 the influence of boundaries is minimized by prescribing homogeneous Neumann boundary condition for g at both walls. Simulations begin in a flowing state at μw = 0.35, then μw is progressively decreased to 0.25 before being ramped back up to 0.35 in order to induce flow arrest and restart. We ensure that the ramp rate is slow enough that it does not affect the results. Most importantly, we perform these simulations for various values of the scalar parameter A prescribing the strength of nonlocal effects, so that we can pinpoint the specific contributions of nonlocal diffusion and non-monotonicity of the limiting local rheology (8) to each of the three features (F1–F3).

Fig. 4(a)–(c) display the time-dependent dimensionless velocity at the top wall, w(t), versus the applied stress ratio, μw(t), for (a) A = 0, (b) A = 0.03, and (c) A = 0.9. Here, the dimensionless velocity is defined as image file: d1sm00659b-t23.tif. The down stress ramp is shown in blue while the up stress ramp is shown in black, as depicted by the arrows. Further, Fig. 4(d)–(f) display the g fields corresponding to the two states indicated by the lone circle and cross on the down and up ramps at μw = 0.33. Correspondingly, these fields are shown with circle or cross markers depending on the stress ramp that they belong to, and they are superimposed to the stable (green) and unstable (red) steady-state solutions gloc of the local fluidity eqn (7) under the same stress ratio and pressure fields. A movie version of Fig. 4, which follows the state of the system as the applied stress ratio μw(t) is progressively decreased and increased, is also included in the ESI. We begin with the case A = 0, for which nonlocal effects are turned off and hence the fluidity eqn (5b) is identical with its local limit (7). As shown in Fig. 4(a), the non-monotonicity of the limiting local rheology (8) leads to different μwversus vw branches in Fig. 4(a) when the stress is ramped down or up. Indeed, the bistable behavior of (7) for image file: d1sm00659b-t24.tif implies that there are two stable steady-state solutions gloc – one flowing and one arrested – within a range of heights, as shown by the green branches in Fig. 4(d). When the applied stress is ramped down, the bistable region moves towards smaller (shallower) values of z, that were previously flowing; thus, the time-dependent solution for g will remain on the flowing branch. Conversely, when the applied stress is ramped up, the bistable region moves towards larger (deeper) values of z, that were previously arrested; thus, the time-dependent solution will remain on the arrested branch. This explains why the down ramp flows at a higher wall velocity vw than the up ramp in Fig. 4(a), which also causes flow to arrest at a lower wall stress ratio, μ*, than that at which it restarts, image file: d1sm00659b-t25.tif. Further, the onset of flow during the up ramp is characterized by a smooth increase in the top wall velocity since the thickness of the flowing layer beneath the top wall smoothly increases from zero (see movie). In conclusion, the non-monotonicity of the local rheology suffices to endow the flow-arrest transition with feature (F1), but features (F2) and (F3) are absent – there is no finite velocity jump at flow onset, and the starting and stopping stresses are identical with the local rheology predictions in Fig. 3(c).


image file: d1sm00659b-f4.tif
Fig. 4 Qualitative behavior of the NGF model with hysteresis in stress-driven plane shear with gravity. (a–c) Time-dependent dimensionless velocity at the top wall, w(t), versus the applied stress ratio at the wall, μw(t). (d–f) Fluidity field g(z,t) corresponding to the state indicated in (a–c) by the lone circle (cross) on the down (up) branch at μw = 0.33. The field belonging to the down (up) branch is shown with circle (cross) markers and is superimposed to the stable (green) and unstable (red) steady-state solutions gloc of the local fluidity eqn (7) under the stress ratio field (12). A movie version of this figure, which follows the state of the system as μw(t) is progressively decreased and increased, is also included in the ESI.

We then turn to the case A = 0.03, corresponding to a tiny amount of nonlocal effects. Fig. 4(b) suggests that in this situation, the main role of the nonlocal diffusion term in (5b) is to merge the flowing section of the up stress ramp with that of the down stress ramp, which is almost unchanged from the case A = 0. In other words, there is only one possible flowing solution for every value of μw, and Fig. 4(e) shows that in the bistable range of heights, this unique solution follows the flowing local solution gloc and not the arrested one. This is a direct consequence of the regularizing effect of the diffusion term, which acts to minimize discontinuities in the fluidity profile. A crucial side effect is that a finite velocity jump emerges at flow onset, since the entire bistable region beneath the top wall suddenly jumps from the arrested to the flowing local solution (see movie). The interplay between nonlocality and non-monotonicity of the local rheology is therefore critical in achieving feature (F2), with (F3) the only one that remains unaccounted for.

Finally, we investigate the case A = 0.9, corresponding to the real calibrated value that we use later. Fig. 4(b) shows that similar to the case A = 0.03, there is only one possible flowing solution for every value of μw. However, the increased strength of nonlocal diffusion leads to a different μwversus vw relationship than before, with much higher wall stress ratios at flow arrest and onset. This is caused by the diffusion term spreading fluidity towards the μ < μ* region where the local solution is arrested, as revealed in Fig. 4(f). Thus, when μw is hardly higher than the stopping and starting stress ratios observed in the case A = 0.03, the μ < μ* region acts as a fluidity sink that prevents the overall nonlocal solution from flowing. The resulting strengthening of the wall stress ratio at flow onset and arrest is dependent on the degree of nonuniformity of the stress ratio (12) controlled by the loading length scale [small script l], and increases with decreasing [small script l]. As a result, nonlocal diffusion induces strengthening of the threshold for flow onset and arrest with reducing system size, which generalizes a similar conclusion from previous studies31,38 that only looked at the stopping stress.

To summarize, non-monotonicity and nonlocality are seen to contribute in different ways to the features (F1–F3) of the flow arrest transition in the case of plane shear with gravity:

(F1) hysteresis of the critical stresses for flow onset and arrest is achieved through non-monotonicity of the local rheology;

(F2) the finite velocity jump at flow onset requires an interplay between non-monotonicity and nonlocal diffusion;

(F3) increased strengthening at smaller system sizes is caused by nonlocal diffusion.

Clearly, all three features are simultaneously achievable only when both non-monotonicity and nonlocality are included in the model. Even though these results have been obtained in a plane shear under gravity configuration, they should hold in other geometries that display a similar spatially-varying stress ratio profile such as annular shear between concentric cylinders, since the mechanisms at play are similar.32,43 Finally, the conclusions that we have reached should apply to other nonlocal rheological models, including in particular those that treat the inertial number I as an order parameter in place of g.42,44

3 Comparisons with DEM simulations

Now that we have established that the NGF model with hysteresis is able to reproduce qualitatively the various features of the flow-arrest transition, the next step is to compare quantitatively model predictions with discrete element method (DEM) simulations in a variety of geometries. To do so, we first calibrate the rheological parameters of the model using DEM simulations of homogeneous, simple plane shear. The calibrated model is then compared with DEM simulations in plane shear under gravity and inclined plane geometries.

We first describe the general setup of our DEM simulations, which we perform in the open-source software LAMMPS.45 We simulate 2D disks with mean diameter d = 0.0008 m and aerial density ρs = 1.3 kg m−2, corresponding to a characteristic grain mass m = ρsπd2/4. The disk diameters are uniformly distributed within ±20% of d in order to prevent crystallisation. Following seminal previous work,11 we use the standard spring-dashpot model with Coulomb friction for the contact force between overlapping particles.46 More specifically, the normal force is given by image file: d1sm00659b-t26.tif where δn ≥ 0 is the normal contact overlap, kn the normal stiffness and gn the damping coefficient, which can be expressed in terms of the coefficient of restitution e for a binary collision as image file: d1sm00659b-t27.tif. The tangential force is given by Ft = ktδt where δt is the accumulated tangential contact displacement and kt is the tangential stiffness, and its magnitude is limited by the surface friction coefficient μsurf so that |Ft| ≤ μsurf|Fn|. Thus, the contact force model is fully described by the parameters kn, kt, e, and μsurf, to which we assign the same values as in past studies.22,32,47 Namely, we use μsurf = 0.4 and we choose kn so that κ = kn/P > 104, with P the characteristic confining pressure, corresponding to the stiff grain limit.11,48 Further, we set kt/kn = 1/2 and e = 0.1, with both having little influence on the phenomenology of dense flows of stiff disks.49,50 Finally, we choose a time step equal to 0.1 times the binary collision time image file: d1sm00659b-t28.tif. At the end of each simulation, the particle-wise quantities in each saved system snapshot are coarse-grained into continuum fields according to the procedure described in Appendix A. Because the geometries that we investigate are homogeneous along the x-direction, this spatial averaging procedure returns an instantaneous streamwise velocity field v(z,t), as well as instantaneous stress field components σxx(z,t), σzz(z,t), and σxz(z,t).

3.1 Calibration with simple plane shear

We first simulate a simple plane shear geometry in DEM, since the homogeneity of the flow in this configuration enables us to calibrate the local part of the NGF model, given by the limiting local rheology (8). The configuration is pictured in Fig. 2(a), and consists of two parallel walls of length L = 120d aligned along the horizontal x-direction, and separated by a distance H along the vertical z-direction. The walls consist of glued disks, colored in brown in Fig. 2(a), and they shear a dense collection of enclosed disks, colored according to their relative velocity magnitude in a particular snapshot of a flowing state. Periodic boundary conditions are applied along the x-direction, and the absence of gravity leads to a uniform stress ratio throughout, which is imparted by the walls. The top wall is assigned a horizontal velocity vw that is either directly prescribed, corresponding to a velocity boundary condition, or calculated following a control scheme that simulates an applied tangential stress τw to the top wall through the feedback law [v with combining dot above]w = (τwσxz(z = 0,t))L/Mw, corresponding to a stress boundary condition. Here, the instantaneous tangential stress σxz exerted by the flowing grains is directly evaluated at the wall, and the effective wall mass Mw acts as a damping parameter, which we take as Mw = 2000m. Although velocity-driven DEM simulations of plane shear flow are the norm,7 such simulations miss important features of the flow-arrest transition.51 To our knowledge, stress-driven plane shear simulations have only been implemented in very few studies either through a solid wall,52,53 which corresponds to our setup, or through shearing of the periodic boundaries.54–57 Finally, the pressure at the top wall is maintained close to a target value Pw through a widely used feedback law11 according to which the distance H between the walls evolves as = (Pw + σzz(z = 0,t))L/gw, where the instantaneous normal stress σzz exerted by the flowing grains is directly evaluated at the wall, and gw is a damping parameter that we take as gw = 100(mkn)1/2.

We begin by performing velocity-driven DEM simulations under various prescribed values of the top-wall velocity vw and for two nominal system heights H = 50d and 25d. After each simulation has reached a steady state, we save 4000 system snapshots evenly distributed in time over a minimum additional top-wall shear displacement of 78H. The instantaneous continuum velocity and stress fields produced by the coarse-graining procedure are then averaged in time. As expected from numerous previous studies,7,11 the strain rate and the stress components are all approximately constant in the central part of the sheared layer, four grain diameters away from the walls. We therefore spatially average these quantities into the strain rate [small gamma, Greek, dot above] = 〈|dv/dz|〉, shear stress τ = 〈σxz〉 and pressure P = 〈|σzz|〉, from which we calculate the stress ratio μ = τ/P and the inertial number image file: d1sm00659b-t29.tif. Different values of the prescribed top-wall velocity produce different (μ,I) pairs, which are displayed in Fig. 5(a) for the two nominal system heights considered. We also plot corresponding results from the DEM simulations of Liu and Henann,32 obtained under identical system parameters and for a nominal height H = 60d. The agreement between the two sets of data validates our simulations; furthermore, the negligible difference between the μ(I) curves pertaining to different system heights demonstrates the negligible influence of the walls. Importantly, we observe the presence of a strain-rate weakening regime at low enough values of I, directly corroborating previous simulation results from DeGiuli and Wyart.17 However, a decreasing relationship between shear stress and strain rate is mechanically unstable and typically results in the formation of shear bands that, in turn, render impossible the accurate measurement of the true strain rate in the strain-rate weakening regime.58 Furthermore, past theoretical studies59,60 suggest that in this regime, systems with a nonlocal flow rule select a specific stress state independent of the nominal strain rate imparted by the walls. The NGF model, therefore, cannot be simply calibrated on velocity-driven DEM data if it is to accurately predict onset and arrest of flow under variations of the applied stress.


image file: d1sm00659b-f5.tif
Fig. 5 Calibration of the local part of the NGF model using simulations of plane shear without gravity. (a) Local μ(I) rheology obtained from velocity-driven DEM simulations for two system sizes, compared with data from Liu and Henann.32 (b) Applied stress ratio at the top wall, μw, versus wall-based inertial number, Iw, obtained from stress-driven DEM simulations for H = 50d, in which the stress ratio applied to the upper wall is first slowly ramped down (black lines), then ramped back up (red lines). Different lines correspond to different initial conditions, realized by letting the system flow at μ = 0.4 for varying amounts of time. (c) Determination of the local yield stress ratio image file: d1sm00659b-t30.tif from the stochastic μstart values pertaining to the different realizations. (d) The limiting non-monotonic local rheology (8) is fit in two steps. First, the parameters shared with the monotonic form (4) are calibrated using the velocity-driven DEM data for I ≥ 10−2 (filled circles), producing the monotonic μloc(I) fit. Then, the parameters of the strain-rate weakening term χ(I;κ) are chosen so that image file: d1sm00659b-t31.tif is equal to the value extracted from the stress-driven DEM data (filled square) and the minimum of μ occurs for 10−3 < I* < 10−2, finally producing the non-monotonic μloc(I) fit.

In order to extract the true critical stresses delineating the transition between static and flowing regimes, we run stress-driven DEM simulations of flow onset and arrest under slowly decreasing and increasing ramps of the top-wall stress ratio μw, for a nominal system height H = 50d. Specifically, we assign a time-dependent top-wall shear stress τw(t) = μw(t)Pw according to the control procedure described above, where μw(t) is the target stress ratio applied to the top wall and Pw is the constant target pressure. We run 20 different simulations by letting the system flow at μw = 0.4 for a varying amount of time after it has reached steady state, effectively imparting a different initial microstructure to each simulation. The applied stress ratio μw is then linearly decreased from 0.4 to 0.25 over a time duration of 6.5 × 107τc ≃ 113 s, inducing jamming of the grains. We then let the contact forces relax by decreasing μw from 0.25 to 0 over a time duration of 2 × 107τc ≃ 35 s and keeping μw at 0 for 1 × 107τc ≃ 17 s. Finally, μw is linearly ramped back up, first from 0 to 0.25 over a time duration of 2 × 107τc ≃ 35 s, then from 0.25 to 0.4 over 6.5 × 107τc ≃ 113 s, triggering onset of flow. The ramp rate is slow enough that the system can be assumed to undergo quasi-steady motion. We save 4000 system snapshots evenly distributed in time, from start to end of the stress ramp. At the end of each simulation, we calculate an instantaneous wall-based inertial number image file: d1sm00659b-t32.tif, where vw(t) is a moving time window average over 50 snapshots of the instantaneous wall velocity to smooth out small fluctuations, and Hw is the average true vertical distance between both walls. Thanks to the little amount of observed wall slip|| and the quasi-steady conditions, we may consider both μw and Iw as smooth approximations of the highly-fluctuating instantaneous values of μ and I in the bulk. The resulting μwversus Iw curves are shown in Fig. 5(b) in black and red for the decreasing and increasing stress ramps, respectively, with different curves corresponding to different simulations. Observe the similarity between these curves and the ones shown in Fig. 2(a), with hysteresis (feature F1) and a velocity jump at flow onset (feature F2) clearly visible in both geometries. Besides, our simulations reveal that the critical stress ratio at flow onset is stochastic and noticeably higher than the threshold obtained from velocity-driven simulations in Fig. 5(a) as I vanishes. Such variability in the transition between arrested and flowing states has been observed previously,61 and may be explained by the role played by the specific structure of the particle contact network.3,62,63 On the other hand, the critical stress ratio at flow arrest is much more narrowly distributed and similar to the velocity-driven flow threshold.

Thus, the NGF model needs to be calibrated using data from both velocity-driven and stress-driven DEM simulations, so as to correctly predict the characteristics of both the flowing regime and the transition between arrested and flowing states. However, the critical stress ratio for flow onset observed in the stress-driven simulations is highly stochastic, while the limiting local rheology (8) of the model predicts a deterministic value image file: d1sm00659b-t33.tif. To reconcile this apparent contradiction, we note that because continuum models in general aim to reproduce the ensemble-average behavior of the discrete system across all possible realizations, we expect our NGF model to predict onset of flow so long as any measurable region of the space of ensembles initiates flow. With this in mind, we therefore reduce the stochastic critical stress from DEM to a unique deterministic value that represents the lowest achievable starting stress of the system as follows. For each realization, we first define μstart as the observed μw when Iw last exceeds 10−3 during stress increase. We then assume that the ensemble of stochastic μstart values follows a uniform probability distribution over a finite range (bounded from below by the lowest achievable starting stress), which allows us to fit a linear function to their cumulative distribution (CDF), shown in Fig. 5(c). An estimate for the lowest achievable μstart is given by the x-intercept of the fitted CDF, which we thus assign to image file: d1sm00659b-t34.tif. Finally, we show in Appendix C that the distribution of stochastic μstart values barely changes for smaller nominal heights H = 25d and 10d, which supports our methodology of calculating the local yield stress ratio image file: d1sm00659b-t35.tif predicted by the model using stress-driven simulations at H = 50d.

The parameters of the limiting local rheology (8) can now be calibrated following a two-step approach pictured in Fig. 5(d). First, the parameters shared with the monotonic form (5) are calibrated using the velocity-driven DEM data corresponding to H = 50d and I ≥ 10−2 (filled circles), producing the monotonic μloc(I) fit. The velocity-driven DEM data for I < 10−2 (open circles) is also shown for reference, but is not used in the calibration. Second, the parameters of the strain-rate weakening term χ(I;κ) are chosen so that image file: d1sm00659b-t36.tif is equal to the value extracted from the stress-driven DEM data (filled square) and the minimum of μ occurs for 10−3 < I* < 10−2, finally producing the non-monotonic μloc(I) fit. The resulting parameter values are μs = 0.2610, μ2 = 0.9784, b = 1.6406, a = 0.0116, c = 50, and n = 1/4. We note that once μs and μ2 are known from the first step, a is obtained in the second step from image file: d1sm00659b-t37.tif by inverting (10). In the second step, we have selected the value 1/4 from DeGiuli and Wyart17 for the parameter n that controls the grain stiffness-dependence of the hysteresis amplitude. We have nonetheless verified that choosing instead n = 0, which removes the stiffness dependence, and recalibrating c accordingly produces negligible changes in our results to follow. Finally, we adopt the value A = 0.9 from Liu and Henann32 for the nonlocal amplitude, which they calibrated on DEM data obtained with the same particle and contact force law properties.

3.2 Plane shear under gravity

We now compare the predictions of the calibrated NGF model with stress-driven DEM simulations of plane shear under gravity shown in Fig. 2(b). We have already investigated in Section 2.3 the qualitative behavior of the model in this geometry, in which the gravity field imparts a nonuniform distribution of stress ratio μ(z) characterized by a loading length scale [small script l]; see eqn (12). Besides the presence of gravity, the DEM setup of the system is identical to that of the previous section, except for the nominal distance H = 60d between the walls. The shear stress and pressure at the top wall are controlled according to the feedback schemes described in the previous section, and the bottom wall is fixed.

Similar to the case of simple plane shear, we perform DEM simulations of arrest and onset of flow under decreasing then increasing ramps of applied stress. Specifically, a time-varying top-wall stress ratio μw(t) is prescribed through a time-dependent target shear stress τw(t) = μw(t)Pw and constant target pressure Pw. The applied stress ratio μw(t) follows the same protocol as in Section 3.1, with the only difference being that we start from, and end at, a top value of μw = 0.45 instead of μw = 0.4. Correspondingly, the duration of the decreasing and increasing sections of the stress ramp between 0.45 and 0.25 is lengthened to 9 × 107τc ≃ 157 s so that the rate of change of μw(t) is unaffected. As before, for each length scale [small script l] we run 20 different simulations corresponding to different initial microstructures, by letting the system spend a varying amount of time in the initial shearing period at μw = 0.45. During that period, gravity is first turned off to ensure homogeneous mixing and shearing of the grains, before being turned back on. We save 4000 system snapshots for each simulation, from which we calculate vw(t), a moving time window average over 50 snapshots of the instantaneous wall velocity. We then calculate the nondimensional wall velocity as image file: d1sm00659b-t38.tif. In order to compare deterministic predictions from the NGF model with the DEM results corresponding to all N = 20 different runs, we transform the discrete values (i)w(t) for i = 1,…,N at each time step into a continuous probability density function (PDF) as follows:

 
image file: d1sm00659b-t39.tif(13)
where δ is the Dirac delta function, and w is the Gaussian kernel
 
image file: d1sm00659b-t40.tif(14)
with L the kernel size, which we choose equal to 0.01 times the maximum observed value of (i)w(t) across all runs and times. The PDF (13) can be conveniently expressed as
 
image file: d1sm00659b-t41.tif(15)
and it integrates to one, as it should. Further, we also need to reduce the stochastic transition stresses between flowing and arrested states into a unique deterministic value, which we define in a similar way to Section 3.1 as the lowest achievable value to be consistent with the methodology followed in calibrating the local yield stress ratio image file: d1sm00659b-t42.tif of the model. For each realization, we first calculate μw,start as the observed μw when w last exceeds the threshold value 10−3 during stress increase, and μw,stop as the observed μw when w last falls below the same threshold during stress decrease. The deterministic starting and stopping critical stress ratios are then defined as the x-intercept of a fitted linear CDF to the stochastic μw,start and μw,stop values, similar to Fig. 5(c). In the following, we will refer to these deterministic thresholds as μw,start and μw,stop.

We also compute NGF model predictions in the same geometry, using the calibrated parameter values from Section 3.1. As described in Section 2.3, the fluidity eqn (5b) reduces to a one-dimensional PDE for g(z,t) that requires the stress ratio profile μ(z,t) as input. Thanks to the homogeneous and quasi-steady conditions, the latter is given through (12) and set by the top-wall stress ratio μw(t), to which we assign the exact same temporal protocol as in the DEM simulations. Since the NGF model is not expected to be valid in the vicinity of the walls, we end the corresponding simulation domain at a distance dw = 2d away from the real walls. Furthermore, we follow previous work32 in using homogeneous Neumann (∂g/∂z = 0) boundary conditions on g at the walls in order to minimize their influence. The details of the discretization method for the ODE governing g(z,t) are presented in Appendix B. Once g(z,t) is known, the strain rate [small gamma, Greek, dot above](z,t) can be calculated using the flow rule (5a) and integrated into the velocity profile v(z,t), taking into account a slip length equal to dw for the velocity at the bottom wall. Finally, the top-wall velocity is extrapolated as vw(t) = v(z = −dw,t) + dw[small gamma, Greek, dot above](z = −dw,t) and is nondimensionalized into w(t).

Fig. 6(a), (b), (d) and (e) display the relationship between μw and w obtained from both DEM simulations and NGF model predictions for two different loading length scales of (a and d) [small script l] = 100d and (b and e) [small script l] = 25d. The increasing stress ramp is shown in (a and b) while the decreasing stress ramp is shown in (d and e). The DEM results are displayed as contours of f(w) corresponding to each value of μw, in such a way that the plots can be read as the probability of occurrence of individual realizations, with yellow color indicating high probability and blue color indicating low probability. The deterministic μw,start and μw,stop values from DEM are also displayed as filled circles, and the NGF model prediction is shown as the red line. An excellent agreement between NGF and DEM is observed in the flowing regime.** Similarly, the transition between arrested and flowing states occurs at similar stress levels in both cases, and displays all three features (F1–F3) identified in the introduction: hysteresis between onset and arrest, velocity jump at onset, and strengthening with smaller [small script l]/d. The amount of velocity jump at flow onset exhibited by the NGF solution appears smaller than that observed in the DEM simulations; this is a consequence of the NGF model being calibrated so as to start flowing at the lowest possible critical stress based on the DEM simulations.


image file: d1sm00659b-f6.tif
Fig. 6 Comparison between NGF model predictions and DEM simulations for plane shear under gravity. (a, b, d and e) Relationship between stress ratio μw and dimensionless velocity w at top wall from DEM (contour of probability values extracted from 20 different runs) and NGF (red lines) for two different loading length scales of (a and d) [small script l] = 100d and (b and e) [small script l] = 25d, with the (a and b) increasing and (d and e) decreasing stress ramps shown separately. The filled circles display the deterministic μw,start and μw,stop critical stress ratios obtained from all DEM runs. (c and f) Critical stress ratios μw,start and μw,stopversus dimensionless loading length scale [small script l]/d from DEM (filled circles) and NGF (continuous line). The crosses show the individual transition stresses pertaining to each of the 20 different runs, while the filled circles represent the deterministic value extracted from the linear fit of the CDF.

Finally, Fig. 6(c) and (f) display the critical stresses μw,start and μw,stopversus the dimensionless loading length scale [small script l]/d. Shown are the individual transition stresses from all 20 DEM runs (crosses), the deterministic values extracted from the linear fit of the CDF (filled circles), and the corresponding NGF predictions (continuous line). The NGF predictions of the critical stresses are obtained using the methodology described in Appendix D, which circumvents the need to run time-dependent simulations for every value of [small script l]/d. Overall, the NGF model predicts a similar amount of strengthening as apparent in the DEM simulations. The slightly higher critical stresses exhibited in DEM for large loading length scales is caused by an observed change in the slope of the g field at the boundary as the pressure applied by the top wall increases. Accordingly, implementing a homogeneous Robin boundary condition for g with a finite associated length scale would lead the NGF model to predict higher values for the transition stresses, closing the gap with the DEM data. Lastly, Appendix E shows that the hysteresis amplitude μw,startμw,stop is only weakly dependent on the dimensionless loading length scale [small script l]/d.

3.3 Inclined plane

As a last example, we evaluate predictions from the calibrated NGF model against DEM simulations in the inclined plane configuration shown in Fig. 2(c). A fixed basal wall of length L = 120d and consisting of glued disks is inclined at an angle θ with respect to the horizontal, and is covered by a dense collection of freely moving disks forming a layer of height H. The x- and z-directions are parallel and orthogonal to the base wall, respectively, and periodic boundary conditions are applied along the x-direction. The gravity field imparts a uniform ratio of shear stress to pressure throughout the layer at steady state equal to μ = tan[thin space (1/6-em)]θ, making this configuration inherently stress-driven. Even though μ is uniform, nonlocal effects still arise from the presence of the rough base, which acts as a sink for velocity fluctuations within the moving grains. The transition between onset and arrest of flow on an inclined plane exhibits all three features mentioned in the introduction, as documented in many past experimental and computational studies.4,9,15,49,64–66

Following the previous cases, we run DEM simulations of flow arrest and onset by slowly decreasing then increasing the inclination angle. Specifically, we prescribe a temporal profile for θ(t) such that the stress ratio μ(t) = tan[thin space (1/6-em)]θ(t) follows the same protocol as in Section 3.1, with an initial flowing period at μ = 0.4 followed by a continuous decrease to μ = 0 and a continuous increase back to μ = 0.4. As before, we execute 20 runs for each layer height H, each with a different time duration spent in the initial flowing regime at μ = 0.4, giving a unique microstructure to every simulation before the start of the stress ramp. We save 4000 system snapshots in each simulation, from which we compute the instantaneous continuum velocity field v(z,t). Anticipating that the NGF model will be run over a truncated domain ending at a distance dw = 2d away from the bottom wall and ds = 3d away from the layer's free surface, we then calculate a depth-averaged instantaneous velocity [v with combining macron](t) over the corresponding truncated domain, which we smooth out using a moving time window average over 50 snapshots. We then express [v with combining macron](t) in terms of a dimensionless Froude number defined as image file: d1sm00659b-t43.tif. As was done in the previous section, the discrete values Fr(i)(t) at each time step, for i = 1,…,N corresponding to all N = 20 different runs, are transformed into a continuous PDF f(Fr) through the convolution (13). Finally, the stochastic transition stresses between flowing and arrested states are reduced into deterministic numbers μstart and μstop according to the procedure detailed in Sections 3.1 and 3.2. For each realization, we first calculate μstart as the observed μ when Fr last exceeds the threshold value 10−2 during stress increase, and μstop as the observed μ when Fr last falls below the same threshold during stress decrease. The deterministic starting and stopping critical stress ratios are then defined as the x-intercept of a fitted linear CDF to the stochastic μstart and μstop values, similar to Fig. 5(c). In the following, we will refer to these deterministic thresholds as μstart and μstop.

We compute predictions of the NGF model in the same geometry, using the calibrated parameter values from Section 3.1 and the same temporal protocol for θ(t) as in the DEM simulations. The fluidity eqn (5b) reduces to a one-dimensional PDE for g(z,t) that takes as input the stress ratio profile, which is still related to the inclination angle as μ(t) = tan[thin space (1/6-em)]θ(t) thanks to the quasi-steady conditions. The uniformity of the stress ratio implies that nonlocal effects are mostly imparted by boundaries, making the choice of boundary conditions critical. Similarly to our approach in Section 3.2, the domain for the NGF solution is defined to start at a distance dw = 2d away from the bottom wall due to the lack of validity of the NGF model near the boundary. At that location, the DEM data suggests a Robin-type homogeneous boundary condition for g, with an associated length scale δ that may be sensitive to various factors. We choose to sidestep the exact modeling of the boundary condition by considering the two edge cases of δ = 0 and δ = ∞, corresponding to homogeneous Dirichlet (g = 0) and Neumann (∂g/∂z = 0) boundary conditions, respectively. Regarding the top boundary, the DEM data shows that the strain rate vanishes about 3 grain diameters below the surface, corroborating previous studies.49,66 We thus end the NGF simulation domain at a distance ds = 3d away from the layer's free surface, and we prescribe a homogeneous Neumann (∂g/∂z = 0) boundary condition for g there, following Kamrin and Henann.31 We also assign a finite pressure to the top boundary equal to the weight of the neglected layer of thickness ds, which is approximately equal to P(z = −ds) = 0.8ρsG(2d)cos[thin space (1/6-em)]θ(t) due to the drop in packing fraction near the layer's surface. The details of the discretization method for the ODE governing g(z,t) are presented in Appendix B. Once g(z,t) is known, the strain rate [small gamma, Greek, dot above](z,t) can be computed through the flow rule (5a). Finally, the velocity can be integrated from [small gamma, Greek, dot above](z,t), taking into account a velocity slip length equal to dw at the bottom boundary of the NGF domain, and depth-averaged to produce [v with combining macron](t) and hence Fr(t).

Fig. 7(a), (b), (d) and (e) display the relationship between μ and Fr obtained from both DEM simulations and NGF model predictions for two different layer heights at rest of (a and d) H = 45.5d and (b and e) H = 9d. The increasing stress ramp is shown in (a and b) while the decreasing stress ramp is shown in (d and e). The DEM results are displayed as contours of f(Fr) corresponding to each value of μ, in such a way that the plots can be read as the probability of occurrence of individual realizations, with yellow color indicating high probability and blue color indicating low probability. The deterministic μstart and μstop values from DEM are also displayed as filled circles, and the NGF model predictions pertaining to the two basal boundary conditions for g are shown as the continuous (homogeneous Neumann) and dash-dotted (homogeneous Dirichlet) red lines. The Neumann boundary condition in the NGF model leads to an excellent agreement with DEM in the flowing regime. However, it does not predict any strengthening of the critical transition stresses μstart and μstop for smaller layer height H. The Dirichlet boundary condition, on the other hand, reproduces the strengthening of the transition stresses but fails to match the DEM results in the flowing regime.


image file: d1sm00659b-f7.tif
Fig. 7 Comparison between NGF model predictions and DEM simulations for inclined plane. (a, b, d and e) Relationship between stress ratio μ = tan[thin space (1/6-em)]θ and Froude number Fr from DEM (contour of probability values extracted from 20 different runs) and NGF (red lines) for two different layer heights of (a and d) H = 45.5d and (b and e) H = 9d, with the (a and b) increasing and (d and e) decreasing stress ramps shown separately. The filled circles display the deterministic μstart and μstop critical stress ratios obtained from all DEM runs. (c and f) Critical stress ratios μstart and μstopversus dimensionless layer height H/d from DEM (filled circles) and NGF (blue lines). The crosses show the individual transition stresses pertaining to each of the 20 different runs, while the filled circles represent the deterministic value extracted from the linear fit of the CDF.

The critical stresses μstart and μstop are depicted in greater detail in Fig. 7(c) and (f)versus the dimensionless layer height H/d. Shown are the individual transition stresses from all 20 DEM runs (crosses), the deterministic values extracted from the linear fit of the CDF (filled circles), and the corresponding NGF predictions using homogeneous Neumann (continuous line) or homogeneous Dirichlet (dash-dotted line) boundary conditions. As we did for planar shear with gravity, the NGF predictions of the critical stresses are obtained using the methodology described in Appendix D, which bypasses the need to run time-dependent simulations for every value of H/d. The Neumann boundary condition does not produce any strengthening of the critical stresses, since it kills the principal cause of flow inhomogeneity in this geometry.†† On the other hand, the Dirichlet boundary condition generates a level of strengthening roughly comparable with the DEM data. As was the case for plane shear under gravity, Appendix E shows that the hysteresis amplitude μw,startμw,stop is only weakly dependent on the dimensionless layer height H/d.

In summary, it appears that the critical stresses are best predicted by the NGF model endowed with a homogeneous Dirichlet boundary condition for g, while the flowing regime is most accurate when a homogeneous Neumann boundary condition is used. This dichotomy could stem either from missing ingredients in the NGF model itself or from an inaccurate description of the boundary condition. Regarding the former, a known shortcoming of the current formulation of the model is that it does not produce the widely documented collapse of the Froude number for all layer heights and angles.64,66,67 We have thus modified the fluidity eqn (5b) following the procedure given in Kamrin and Henann31 so that the model collapses the Froude number far away from flow threshold, in the limit χ → 0. We also tried another version of that procedure replacing the quadratic term with a cubic one as in the model of Lee and Yang.42 Yet neither of these modified models performed substantially better, with the no-slip solution still flowing significantly slower than the DEM data. This points to the boundary condition being the main culprit – incidentally, the true length scale δ at the bottom boundary is observed to increase with flow rate in the DEM data. With a velocity-dependent δ that jumps from near zero in the arrested state to a large value in the flowing state, the NGF model could potentially produce correct predictions of both the transition stresses and the flowing regime. A boundary condition of this type could be formulated as a dynamical system governing the evolution of δ in response to relevant driving quantities. As we note in the conclusion, however, formulating accurate and physically-justified fluidity boundary conditions remains a key open issue within NGF modeling, and such an endeavor is relegated to future work.

4 Conclusions

In this paper, we have studied the combined role of strain-rate weakening behavior and nonlocal effects in explaining key features of the hysteretic transition between solid-like and liquid-like behavior in dense granular materials as the applied stress is ramped up and down. These features include the hysteresis of the critical stresses at flow onset and arrest, the finite jump in velocity during flow onset, and the strengthening of the critical stresses with reducing system size. In a first part, we modified the nonlocal granular fluidity (NGF) model so that it reduces to a non-monotonic local rheology in homogeneous flows. Through numerical simulations of flow onset and arrest in planar shear with gravity using the modified NGF model, we demonstrated qualitatively that the inclusion of both nonlocal effects and non-monotonicity of the local rheology is essential to account for all three features mentioned above.

In a second part, we compared quantitatively predictions of the modified NGF model with DEM simulations of flow onset and arrest in various geometries. First, we calibrated the local parameters of the model using DEM simulations of homogeneous plane shear flow. In so doing, we highlighted the importance of calibrating the local critical stress for flow onset using stress-driven simulations, since measurements based on velocity-driven simulations are unreliable in the strain-rate weakening regime. The stress-driven simulations, however, exhibited large variability in the transition stresses between arrested and flowing regimes. Thus, we developed a criterion to extract a unique deterministic value, corresponding to the lowest observable critical stress, from a large number of repeated runs. We then compared predictions of the calibrated NGF model with stress-driven DEM simulations in planar shear with gravity and inclined plane configurations. In the former case, the model gave accurate predictions of both the transition between flowing and arrested states as well as the characteristics in the flowing regime. In the latter case, the accuracy of the model predictions was strongly affected by the choice of boundary conditions, with no single choice able to reproduce both the transition stresses as well as the average velocity in the flowing regime.

These results suggest that the NGF model generally leads to more accurate predictions when nonlocal effects are mostly generated by the spatial dependence of the μ field, as is the case for planar shear with gravity, rather than by boundaries in the absence of a spatially-dependent μ field, as is the case for inclined plane flow. A possible explanation stems from Liu and Henann's32 observation that an inhomogeneous μ field leads to much stronger size-dependent strengthening than boundary effects, making the accuracy of model predictions less reliant on the particular choice of fluidity boundary conditions when both mechanisms are present. Conversely, model predictions are much more sensitive to the specific type of boundary conditions when the latter are the principal source of nonlocal effects, which calls for better understanding of the interaction between flowing particles and the boundary. In fact, numerous experimental and computational studies have also underscored the sensitivity on wall roughness of transition stresses and velocity profiles in inclined plane flow,65,68–70 plane shear flow without gravity,71 as well as annular shear flow.72 Although progress has been made for flat frictional walls,73,74 the correct modeling of boundary conditions in the general case, as a function of wall properties, is still an open question whose resolution would benefit all nonlocal rheological models.32,41,75

Conflicts of interest

The authors have no conflicts to declare.

A Coarse-graining methodology

In this appendix, we describe our coarse-graining procedure for extracting continuum velocity and stress fields from the particle-wise DEM data. The approach we follow was introduced in Zhang and Kamrin,35 building upon earlier work.2,22,70 Since the geometries that we consider are homogeneous along the x-direction, the spatial averaging generates fields defined at discrete heights zk, spaced 0.5d apart. For a given position zk, we define the instantaneous velocity and stress fields as
 
image file: d1sm00659b-t44.tif(16)
 
image file: d1sm00659b-t45.tif(17)
where [v with combining macron] and [small sigma, Greek, macron] are sublayer-wise velocity and stress averages at the heights zm = zk + (W/2)(m/M), each weighted by the coefficient wm = cos((π/2)(m/n)). Following Zhang and Kamrin,35 we choose W = 2d and M = 5. We now let Lim denote the cross-sectional length between particle i and the horizontal line at height zm. The sublayer-wise instantaneous velocity is given by
 
image file: d1sm00659b-t46.tif(18)
where vi is the velocity of grain i. Then, the sublayer-wise instantaneous stress field is defined by
 
image file: d1sm00659b-t47.tif(19)
where L is the domain length along the x-direction, and σi is the stress tensor associated with grain i. The latter consists of contact and kinetic contributions, and is given by
 
image file: d1sm00659b-t48.tif(20)
where Ai = πdi2/4 and mi = ρsAi are respectively the area and mass of grain i, fij is the contact force exerted on grain i by grain j, and rij is the vector pointing from the center of grain i to that of grain j. In the kinetic contribution, the velocity fluctuations are calculated as δvi(t) = vi(t) − v(zi,t), where v(zi,t) is the coarse-grained instantaneous velocity (16) interpolated to the vertical position zi of grain i.

B Numerical discretization

Here, we present our numerical discretization method for solving the fluidity eqn (5b) under a time-dependent applied stress. In the problems that we consider, the equation reduces to a one-dimensional PDE for g(z,t) that is driven by a prescribed stress ratio function μ(z,t). The physics governing the time scale t0 that appears in the fluidity equation are still unknown, and we simply assign a sufficiently small value t0 = 10−4 s that it does not affect the dynamics of the solution. The spatial domain is discretized into N = 100 nodes, and the diffusion term is evaluated using second-order finite differences. The fluidity equation is then integrated in time using an implicit Euler scheme with a time step Δt = 5 × 10−4 s, which we implement in MATLAB. Further, we artificially limit g(z,t) to a minimum value of 10−2 s−1 in order to avoid g(z,t) reaching infinitesimally small values during the arrested portion of the stress ramp. Indeed, this would prevent g(z,t) from growing sufficiently fast when flow onset should occur, as the applied stress is subsequently increased. We have verified that the floor value for g(z,t) is small enough that it does not alter the observed transition stresses.

C Size effects in simple plane shear

We report in Fig. 8 additional DEM results on the system-size dependence of the critical transition stresses μstart and μstop in plane shear without gravity. Mirroring our definition of μstart in Section 3.1, μstop is defined as the observed μw when Iw last falls below 10−3 during stress decrease. Shown are the individual transition stresses from 20 DEM runs (crosses) and the corresponding deterministic values (filled circles) extracted from the linear fit of the CDF according to the procedure outlined in Fig. 5(c). We observe that the critical stresses are almost independent of system size, corroborating results from a previous DEM study.75
image file: d1sm00659b-f8.tif
Fig. 8 Critical stresses μstart and μstop in simple plane shear versus true system height H/d, obtained from 20 different DEM runs. The crosses show the individual transition stresses pertaining to every run, while the filled circles represent the deterministic value extracted from the linear fit of the CDF.

D Critical stresses from NGF

In this appendix, we explain how to obtain the critical starting and stopping stresses predicted by the NGF model without computing time-dependent solutions to a slowly varying applied stress, which are computationally intensive due to the required low rate of change of the applied stress to ensure quasi-steady conditions. Let the scalar [small mu, Greek, macron] denote the amplitude of the applied stress ratio μ(z) throughout the domain – for instance, [small mu, Greek, macron] is μw in the case of plane shear with gravity, or tan[thin space (1/6-em)]θ in inclined plane. For a given geometry, implying a certain distribution of the stress ratio μ(z), we then rewrite the fluidity eqn (5b) as
 
t0ġ = [scr F, script letter F](g;[small mu, Greek, macron]),(21)
where the dependence on the magnitude of μ(z) has been explicitly denoted through [small mu, Greek, macron]. In the following, we will call g0 any steady-state solution of (21).

At low [small mu, Greek, macron], the arrested state g0 = 0 is the only stable solution. Gradually increasing [small mu, Greek, macron], flow onset occurs the moment the g0 = 0 solution becomes unstable to small perturbations g′, which defines the critical starting stress [small mu, Greek, macron]start.41 These small perturbations are governed by the linear equation

 
image file: d1sm00659b-t49.tif(22)
where [script L](g0;[small mu, Greek, macron]) – the linearization (also called Fréchet derivative) of [script L](g;[small mu, Greek, macron]) around g0 – acts on the perturbation g′ as
 
image file: d1sm00659b-t50.tif(23)
To evaluate whether perturbations grow or decay, we substitute image file: d1sm00659b-t51.tif into (22), which leads to the eigenvalue problem
 
image file: d1sm00659b-t52.tif(24)
for the growth rate λ. This eigenvalue problem can be solved numerically by discretizing [script L](g0;[small mu, Greek, macron]) using a finite difference approximation, giving a spectrum of eigenvalues with the one having the largest real part, λm, dictating the overall rate of growth or decay of the perturbation. Setting g0 = 0, one can perform repeatedly this calculation for increasing values of [small mu, Greek, macron] until Re{λm} becomes positive, at which point the arrested solution loses stability and [small mu, Greek, macron] = [small mu, Greek, macron]start.

Gradually decreasing [small mu, Greek, macron] from a value above [small mu, Greek, macron]start, flow arrest occurs the moment (21) ceases to admit a nonzero steady-state solution g0, which defines the critical stopping stress [small mu, Greek, macron]stop. To check whether that is true at a given value of [small mu, Greek, macron], it suffices to perform Newton–Raphson iterations to find g0. Starting from an initial guess g0, the algorithm performs at each step n the update gn+1 = gn + ωΔn, where 0 < ω < 1 is a relaxation parameter and the step direction Δn is given through the linear system

 
[script L](gn;[small mu, Greek, macron])Δn = −[scr F, script letter F](gn;[small mu, Greek, macron]).(25)
We stop the iterations when the norm of [scr F, script letter F](gn;[small mu, Greek, macron]) falls under a specified threshold, indicating that gn has converged to a steady-state solution g0 of (21). Thus, our strategy to find [small mu, Greek, macron]stop goes as follows. We start with a value of [small mu, Greek, macron] above [small mu, Greek, macron]start, for which we are guaranteed a nonzero g0 solution. We compute the latter by letting g reach steady-state in the time-dependent solver. Then, we repeatedly compute g0 for incrementally decreasing values of [small mu, Greek, macron] through Newton–Raphson iterations, using at each step level of [small mu, Greek, macron] the converged solution g0 from the previous step as an initial guess. At some point the Newton–Raphson iterations will suddenly converge to the arrested g0 = 0 solution, indicating that [small mu, Greek, macron] has reached [small mu, Greek, macron]stop.

Both the eigenvalue problem (24) and linear system (25) are implemented in MATLAB borrowing the same grid and discretized differential operators used in the time-dependent solver. In the Newton–Raphson iterations, we use ω = 0.5 to balance stability and speed of convergence. Previous studies31,32 have shown that for some geometries and boundary conditions, there exist analytical or semi-analytical solutions for the growth rate λ and thus the threshold μstart. However, such solutions are much harder to obtain for μstop, despite partial progress in that direction on a I-gradient model applied to the inclined plane scenario.42 Therefore, we limit ourselves in this paper to the numerical methodology that we have outlined above, noting that it is computationally very efficient – the starting and stopping curves in Fig. 6 and 7 were calculated in a few minutes on a laptop.

E Hysteresis size dependence

Here, we investigate the system-size dependence of the hysteresis amplitude, measured by the difference of the starting and stopping critical stress ratios, based on the DEM simulations and NGF model predictions reported in Fig. 6 and 7 in Sections 3.2 and 3.3. Fig. 9 displays μstartμstop as a function of the dimensionless loading length scale [small script l]/d for plane shear with gravity data, and as a function of the dimensionless layer height H/d for inclined plane data. The DEM and NGF data are within a comparable range and suggest a weak effect of system size on the hysteresis amplitude.
image file: d1sm00659b-f9.tif
Fig. 9 Critical stress difference μstartμstopversus dimensionless loading length scale [small script l]/d for plane shear with gravity, and versus dimensionless layer height H/d for inclined plane. The DEM data (filled circles) and NGF data (blue lines) is the same as in Fig. 6(c, f) and 7(c, f).

Notes and references

  1. Y. Forterre and O. Pouliquen, Annu. Rev. Fluid Mech., 2008, 40, 1–24 CrossRef.
  2. B. Andreotti, Y. Forterre and O. Pouliquen, Granular media: between fluid and solid, Cambridge University Press, 2013 Search PubMed.
  3. I. Srivastava, L. E. Silbert, J. B. Lechman and G. S. Grest, 2021, arXiv preprint arXiv:2104.00787.
  4. A. Daerr and S. Douady, Nature, 1999, 399, 241 CrossRef CAS.
  5. A. Lucas, A. Mangeney and J. P. Ampuero, Nat. Commun., 2014, 5, 1–9 Search PubMed.
  6. C. H. Scholz, Nature, 1998, 391, 37–42 CrossRef CAS.
  7. G. MiDi, Eur. Phys. J. E: Soft Matter Biol. Phys., 2004, 14, 341–365 CrossRef CAS PubMed.
  8. F. Da Cruz, F. Chevoir, D. Bonn and P. Coussot, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2002, 66, 051305 CrossRef CAS PubMed.
  9. O. Pouliquen and Y. Forterre, J. Fluid Mech., 2002, 453, 133–151 CrossRef CAS.
  10. S. Courrech du Pont, PhD thesis, Université Paris XI, 2003.
  11. F. Da Cruz, S. Emam, M. Prochnow, J.-N. Roux and F. Chevoir, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2005, 72, 021309 CrossRef PubMed.
  12. J. A. Dijksman, G. H. Wortel, L. T. van Dellen, O. Dauchot and M. van Hecke, Phys. Rev. Lett., 2011, 107, 108303 CrossRef PubMed.
  13. O. Kuwano, R. Ando and T. Hatano, Geophys. Res. Lett., 2013, 40, 1295–1299 CrossRef.
  14. H. Perrin, C. Clavaud, M. Wyart, B. Metzger and Y. Forterre, Phys. Rev. X, 2019, 9, 031027 CAS.
  15. A. Russell, C. Johnson, A. Edwards, S. Viroulet, F. Rocha and J. Gray, J. Fluid Mech., 2019, 869, 313–340 CrossRef CAS.
  16. F.-L. Yang and Y.-T. Huang, Granular Matter, 2016, 18, 1–13 CrossRef.
  17. E. DeGiuli and M. Wyart, Proc. Natl. Acad. Sci. U. S. A., 2017, 114, 9284–9289 CrossRef CAS PubMed.
  18. L. Quartier, B. Andreotti, S. Douady and A. Daerr, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2000, 62, 8299 CrossRef CAS PubMed.
  19. S. Courrech du Pont, P. Gondret, B. Perrin and M. Rabaud, Phys. Rev. Lett., 2003, 90, 044301 CrossRef PubMed.
  20. H. Jaeger, C.-H. Liu, S. Nagel and T. Witten, EPL, 1990, 11, 619 CrossRef.
  21. P. Mills, P. Rognon and F. Chevoir, EPL, 2008, 81, 64005 CrossRef.
  22. G. Koval, J.-N. Roux, A. Corfdir and F. Chevoir, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2009, 79, 021306 CrossRef PubMed.
  23. Z. Tang, T. A. Brzinski, M. Shearer and K. E. Daniels, Soft Matter, 2018, 14, 3040–3048 RSC.
  24. H. J. Melosh, J. Geophys. Res.: Solid Earth, 1979, 84, 7513–7520 CrossRef.
  25. J. Goyon, A. Colin, G. Ovarlez, A. Ajdari and L. Bocquet, Nature, 2008, 454, 84–87 CrossRef CAS PubMed.
  26. K. Reddy, Y. Forterre and O. Pouliquen, Phys. Rev. Lett., 2011, 106, 108301 CrossRef CAS PubMed.
  27. J. Gaume, G. Chambon and M. Naaim, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2011, 84, 051304 CrossRef PubMed.
  28. K. Kamrin, Front. Phys., 2019, 7, 116 CrossRef.
  29. K. Kamrin and G. Koval, Phys. Rev. Lett., 2012, 108, 178301 CrossRef PubMed.
  30. D. L. Henann and K. Kamrin, Proc. Natl. Acad. Sci. U. S. A., 2013, 110, 6730–6735 CrossRef CAS PubMed.
  31. K. Kamrin and D. L. Henann, Soft Matter, 2015, 11, 179–185 RSC.
  32. D. Liu and D. L. Henann, Soft Matter, 2018, 14, 5294–5305 RSC.
  33. G. Picard, A. Ajdari, F. Lequeux and L. Bocquet, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2005, 71, 010501 CrossRef PubMed.
  34. L. Bocquet, A. Colin and A. Ajdari, Phys. Rev. Lett., 2009, 103, 036001 CrossRef PubMed.
  35. Q. Zhang and K. Kamrin, Phys. Rev. Lett., 2017, 118, 058001 CrossRef PubMed.
  36. A. Bhateja and D. V. Khakhar, Phys. Rev. Fluids, 2018, 3, 062301 CrossRef.
  37. S. Kim and K. Kamrin, Phys. Rev. Lett., 2020, 125, 088002 CrossRef CAS PubMed.
  38. D. Liu and D. L. Henann, J. Fluid Mech., 2017, 831, 212–227 CrossRef CAS.
  39. D. L. Henann and K. Kamrin, Phys. Rev. Lett., 2014, 113, 178001 CrossRef PubMed.
  40. S. Li and D. L. Henann, Phys. Rev. E, 2020, 102, 022908 CrossRef CAS PubMed.
  41. I. S. Aranson and L. S. Tsimring, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2002, 65, 061303 CrossRef PubMed.
  42. K.-L. Lee and F.-L. Yang, Phys. Rev. E, 2017, 96, 062909 CrossRef PubMed.
  43. O. Pouliquen and Y. Forterre, Philos. Trans. R. Soc., A, 2009, 367, 5091–5107 CrossRef PubMed.
  44. M. Bouzid, M. Trulsson, P. Claudin, E. Clément and B. Andreotti, Phys. Rev. Lett., 2013, 111, 238301 CrossRef PubMed.
  45. S. Plimpton, J. Comput. Phys., 1995, 117, 1–19 CrossRef CAS.
  46. P. A. Cundall and O. D. Strack, Geotechnique, 1979, 29, 47–65 CrossRef.
  47. K. Kamrin and G. Koval, Comput. Part. Mech., 2014, 1, 169–176 CrossRef.
  48. J.-N. Roux and G. Combe, C. R. Phys., 2002, 3, 131–140 CrossRef CAS.
  49. L. E. Silbert, D. Ertas, G. S. Grest, T. C. Halsey, D. Levine and S. J. Plimpton, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2001, 64, 051302 CrossRef CAS PubMed.
  50. C. S. Campbell, J. Fluid Mech., 2002, 465, 261 CrossRef CAS.
  51. I. Srivastava, L. E. Silbert, G. S. Grest and J. B. Lechman, Phys. Rev. Lett., 2019, 122, 048003 CrossRef CAS PubMed.
  52. D. Volfson, L. S. Tsimring and I. S. Aranson, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2003, 68, 021301 CrossRef PubMed.
  53. M. P. Ciamarra, R. Pastore, M. Nicodemi and A. Coniglio, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2011, 84, 041308 CrossRef PubMed.
  54. M. Otsuki and H. Hayakawa, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2011, 83, 051301 CrossRef PubMed.
  55. K. C. Smith, I. Srivastava, T. S. Fisher and M. Alam, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2014, 89, 042203 CrossRef PubMed.
  56. I. Srivastava and T. S. Fisher, Soft Matter, 2017, 13, 3411–3421 RSC.
  57. I. Srivastava, L. E. Silbert, G. S. Grest and J. B. Lechman, J. Fluid Mech., 2021, 907, A18 CrossRef CAS.
  58. T. Divoux, M. A. Fardin, S. Manneville and S. Lerouge, Annu. Rev. Fluid Mech., 2016, 48, 81–103 CrossRef.
  59. X.-F. Yuan, EPL, 1999, 46, 542 CrossRef CAS.
  60. C.-Y. D. Lu, P. D. Olmsted and R. Ball, Phys. Rev. Lett., 2000, 84, 642 CrossRef CAS PubMed.
  61. A. H. Clark, M. D. Shattuck, N. T. Ouellette and C. S. O'Hern, Phys. Rev. E: Stat., Nonlinear, Soft Matter Phys., 2015, 92, 042202 CrossRef PubMed.
  62. D. Bi, J. Zhang, B. Chakraborty and R. P. Behringer, Nature, 2011, 480, 355–358 CrossRef CAS PubMed.
  63. A. H. Clark, J. D. Thompson, M. D. Shattuck, N. T. Ouellette and C. S. O'Hern, Phys. Rev. E, 2018, 97, 062901 CrossRef CAS PubMed.
  64. O. Pouliquen, Phys. Fluids, 1999, 11, 542–548 CrossRef CAS.
  65. P. Rognon, MSc thesis, Université de Marne la Vallée, 2002 Search PubMed.
  66. L. E. Silbert, J. W. Landry and G. S. Grest, Phys. Fluids, 2003, 15, 1–10 CrossRef CAS.
  67. Y. Forterre and O. Pouliquen, J. Fluid Mech., 2003, 486, 21–50 CrossRef.
  68. L. E. Silbert, G. S. Grest, S. J. Plimpton and D. Levine, Phys. Fluids, 2002, 14, 2637–2646 CrossRef CAS.
  69. C. Goujon, N. Thomas and B. Dalloz-Dubrujeaud, Eur. Phys. J. E: Soft Matter Biol. Phys., 2003, 11, 147–157 CrossRef CAS PubMed.
  70. T. Weinhart, R. Hartkamp, A. R. Thornton and S. Luding, Phys. Fluids, 2013, 25, 070605 CrossRef.
  71. P. Schuhmacher, F. Radjai and S. Roux, EPJ Web of Conferences, 2017, p. 03090.
  72. F. Fazelpour and K. E. Daniels, EPJ Web of Conferences, 2021, p. 03014.
  73. R. Artoni, A. C. Santomaso, M. Go and P. Canu, Phys. Rev. Lett., 2012, 108, 238002 CrossRef PubMed.
  74. R. Artoni and P. Richard, Phys. Rev. Lett., 2015, 115, 158001 CrossRef PubMed.
  75. P. Chaudhuri, V. Mansard, A. Colin and L. Bocquet, Phys. Rev. Lett., 2012, 109, 036001 CrossRef PubMed.

Footnotes

Electronic supplementary information (ESI) available. See DOI: 10.1039/d1sm00659b
By system size, we refer to the relevant length scale controlling the width of shear regions in the flow field. Depending on the geometry, this length scale can either be geometric or stress-induced.
§ We note that the qualitative behavior of the model is independent of its specific parameter values. The latter are therefore chosen to be reasonably close to the calibrated values obtained later in Section 3.1, while displaying the hysteretic behavior of the model with enough clarity in Fig. 3 and 4.
Previous DEM simulations32 have shown that the vertical extent of the shear region below the top wall scales with [small script l]. For small enough [small script l], the bottom region is thus quasi-static, and the results are independent of domain height.
|| In our DEM simulations, the relative difference between the upper wall velocity and the coarse-grained streamwise velocity of the grains near the upper wall never exceeds 3%.
** The small discrepancy observed in the case [small script l]/d = 25 is probably attributable to the choice of boundary conditions for g. An extensive discussion on the role of the latter is presented in Section 3.3.
†† In the presence of the Neumann boundary condition, a small amount of flow inhomogeneity is still incurred by the pressure-dependent quadratic term in the fluidity eqn (5b), which explains the slight dependence of μstop on H/d in Fig. 7(f).

This journal is © The Royal Society of Chemistry 2021