+ All Categories
Home > Documents > NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among...

NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among...

Date post: 14-Feb-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
22
REVIEW COPY NOT FOR DISTRIBUTION Micro-Macro Correlations and Anisotropy in Granular Assemblies under Uniaxial Loading and Unloading. Olukayode I. Imole, Mateusz Wojtkowski, Vanessa Magnanimo and Stefan Luding Multi Scale Mechanics (MSM), Faculty of Engineering Technology, MESA+, University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands. (Dated: October 30, 2013) The influence of contact friction on the behavior of dense, polydisperse granular assemblies under uniaxial (oedometric) loading and unloading deformation is studied using discrete element simula- tions. Even though the uniaxial deformation protocol is one of the “simplest” element tests possible, the evolution of the structural anisotropy necessitates its careful analysis and understanding, since it is the source of interesting and unexpected observations. On the macroscopic, homogenized, continuum scale, the deviatoric stress ratio and the deviatoric fabric behave in a different fashion during uniaxial loading and unloading. The maximal stress ratio and strain increase with increasing contact friction. In contrast, the deviatoric fabric reaches its maximum at a unique strain level independent of friction. For unloading, the reversal of stress displays a friction-dependent delay, while the reversal of fabric is found to occur also with delay, but slightly earlier. On the micro-level, a friction-dependent non-symmetry of the proportion of weak/strong and sliding/sticking contacts with respect to the total contacts during loading and unloading is observed. Coupled to this, from the directional probability distribution, the “memory” and history-dependent behavior of granular systems is confirmed. Surprisingly, while a rank-2 tensor is sufficient to describe the evolution of the normal force directions, a sixth order harmonic approximation is necessary to describe the contact probability distribution, the tangential force and the mobilized friction. We conclude that the simple uniaxial deformation activates microscopic phenomena not only in the active Cartesian directions, but also at intermediate orientations, with the tilt angle being dependent on friction, so that the microstructure and forces cause the interesting macroscopic behavior. PACS numbers: 45.70.Cc, 81.05.Rm, 81.20.Ev I. INTRODUCTION AND BACKGROUND Granular materials are omnipresent in nature and widely used in various industries ranging from food, pharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band localization, history- dependence, jamming and yield have attracted significant scientific interest over the past decade [1, 17, 28]. The bulk behavior of these materials depends on the behav- ior of their constituents (particles) interacting through contact forces. To understand their deformation behav- ior, various laboratory element tests can be performed [34, 40]. Element tests are (ideally homogeneous) macro- scopic tests in which one can control the stress and/or strain path. Such macroscopic experiments are impor- tant ingredients in developing and calibrating constitu- tive relations, but they provide little information on the microscopic origin of the bulk flow behavior. An alterna- tive is the Discrete Element Method (DEM) [28], since it provides information about the micro-structure beyond what is experimentally accessible. One element test which can easily be realized (experi- mentally or numerically) is the uniaxial (or oedometric) compression (in a cylindrical or box geometry) involving an axial deformation of a bulk sample while the lateral boundaries of the system are fixed. This test is partic- ularly suited for determining the poroelastic properties of granular materials. During uniaxial loading, isotropic compression and non-isotropic deformation (pure shear) are superposed, so that pressure and shear stress build up. After reversal, pressure and shear stress decay and the latter changes sign after a finite strain, which de- pends on friction. When a granular material is sheared, along with the shear stress, also anisotropy of the contact network begins to develop. It is known that besides density and stress, geometric anisotropy (contact fabric) is an important ingredient to fully understand the micro-macro-mechanics of granular materials. In addition, the effects of contact friction be- tween the constituent grains influences the micromechan- ical response under uniaxial loading, such that a rather simple element test begins to reveal interesting features. Several studies have numerically investigated the extent to which the response of granular media is affected by friction [2–4, 42], especially in the triaxial geometry but not many studies exist on uniaxial loading and unloading of frictional systems [8]. Also, the transmission of stress between grain contacts is relevant, as detailed in this study. Experimental visu- alizations of the distribution of forces using photo-elastic particles in 2D is about the only way to access this infor- mation – see [33, 56] and references therein – even though 3D photoelasticity and other neutron diffraction methods [54] have also been employed. Earlier numerical stud- ies have highlighted the particular character of the con- tact force network, showing that strong contacts carry-
Transcript
Page 1: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

REVIEW

COPY

NOT FOR D

ISTRIB

UTION

Micro-Macro Correlations and Anisotropy in Granular Assemblies under Uniaxial

Loading and Unloading.

Olukayode I. Imole, Mateusz Wojtkowski, Vanessa Magnanimo and Stefan Luding

Multi Scale Mechanics (MSM), Faculty of Engineering Technology, MESA+,

University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands.(Dated: October 30, 2013)

The influence of contact friction on the behavior of dense, polydisperse granular assemblies underuniaxial (oedometric) loading and unloading deformation is studied using discrete element simula-tions. Even though the uniaxial deformation protocol is one of the “simplest” element tests possible,the evolution of the structural anisotropy necessitates its careful analysis and understanding, sinceit is the source of interesting and unexpected observations.

On the macroscopic, homogenized, continuum scale, the deviatoric stress ratio and the deviatoricfabric behave in a different fashion during uniaxial loading and unloading. The maximal stressratio and strain increase with increasing contact friction. In contrast, the deviatoric fabric reachesits maximum at a unique strain level independent of friction. For unloading, the reversal of stressdisplays a friction-dependent delay, while the reversal of fabric is found to occur also with delay,but slightly earlier.

On the micro-level, a friction-dependent non-symmetry of the proportion of weak/strong andsliding/sticking contacts with respect to the total contacts during loading and unloading is observed.Coupled to this, from the directional probability distribution, the “memory” and history-dependentbehavior of granular systems is confirmed. Surprisingly, while a rank-2 tensor is sufficient to describethe evolution of the normal force directions, a sixth order harmonic approximation is necessary todescribe the contact probability distribution, the tangential force and the mobilized friction.

We conclude that the simple uniaxial deformation activates microscopic phenomena not only in theactive Cartesian directions, but also at intermediate orientations, with the tilt angle being dependenton friction, so that the microstructure and forces cause the interesting macroscopic behavior.

PACS numbers: 45.70.Cc, 81.05.Rm, 81.20.Ev

I. INTRODUCTION AND BACKGROUND

Granular materials are omnipresent in nature andwidely used in various industries ranging from food,pharmaceutical, agriculture and mining – among others.In many granular systems, interesting phenomena likedilatancy, anisotropy, shear-band localization, history-dependence, jamming and yield have attracted significantscientific interest over the past decade [1, 17, 28]. Thebulk behavior of these materials depends on the behav-ior of their constituents (particles) interacting throughcontact forces. To understand their deformation behav-ior, various laboratory element tests can be performed[34, 40]. Element tests are (ideally homogeneous) macro-scopic tests in which one can control the stress and/orstrain path. Such macroscopic experiments are impor-tant ingredients in developing and calibrating constitu-tive relations, but they provide little information on themicroscopic origin of the bulk flow behavior. An alterna-tive is the Discrete Element Method (DEM) [28], since itprovides information about the micro-structure beyondwhat is experimentally accessible.

One element test which can easily be realized (experi-mentally or numerically) is the uniaxial (or oedometric)compression (in a cylindrical or box geometry) involvingan axial deformation of a bulk sample while the lateralboundaries of the system are fixed. This test is partic-ularly suited for determining the poroelastic properties

of granular materials. During uniaxial loading, isotropiccompression and non-isotropic deformation (pure shear)are superposed, so that pressure and shear stress buildup. After reversal, pressure and shear stress decay andthe latter changes sign after a finite strain, which de-pends on friction. When a granular material is sheared,along with the shear stress, also anisotropy of the contactnetwork begins to develop.It is known that besides density and stress, geometric

anisotropy (contact fabric) is an important ingredient tofully understand the micro-macro-mechanics of granularmaterials. In addition, the effects of contact friction be-tween the constituent grains influences the micromechan-ical response under uniaxial loading, such that a rathersimple element test begins to reveal interesting features.Several studies have numerically investigated the extentto which the response of granular media is affected byfriction [2–4, 42], especially in the triaxial geometry butnot many studies exist on uniaxial loading and unloadingof frictional systems [8].Also, the transmission of stress between grain contacts

is relevant, as detailed in this study. Experimental visu-alizations of the distribution of forces using photo-elasticparticles in 2D is about the only way to access this infor-mation – see [33, 56] and references therein – even though3D photoelasticity and other neutron diffraction methods[54] have also been employed. Earlier numerical stud-ies have highlighted the particular character of the con-tact force network, showing that strong contacts carry-

Page 2: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

2

ing force larger than the average, are oriented anisotropi-cally, with preferred direction parallel to the axis of com-pression, while those originating from weak contacts areisotropic or have a weak orientation orthogonal to thecompression axis [38]. Another interesting issue is thedistribution and orientation of tangential forces duringthe deformation of dense frictional packings [37, 46, 47].In early, two-dimensional studies on frictional avalanch-ing [37], it has been observed that friction is mobilizedmostly from weak contacts, whereas strong contacts re-sist friction mobilization.In the present study, we use discrete element simula-

tions to investigate and relate the dependencies betweenthe microscopic observations presented hereafter with theevolution of macroscopic quantities such as pressure anddeviatoric stress – and to further extend this to explainthe evolution of the structural/contact and force/stressanisotropies.We first describe the simulation method and model pa-

rameters along with the preparation and test proceduresin II. The definitions of averaged micro-macro quanti-ties including strain, stress and structural anisotropiesare presented in III. Where given, anisotropy refers tonot only the deviatoric stress, but also to the direction-dependence and inhomogeneity of forces, i.e., its micro-scopic origin. Next, we discuss the results of the currentstudy by presenting the evolution of the stress and struc-tural anisotropies during uniaxial loading and unloadingin IVA followed by the magnitude and orientation oftheir respective eigenvalues in IVB. Furthermore, wediscuss friction mobilization in Section IVC followed bythe probability density functions of the normal and tan-gential forces in IVD and the classification of weak andstrong forces in IVD. In Section V, we discuss the po-lar representation of the contact distribution based onthe constant surface and constant bin width method andextract the structural anisotropy parameters using a 6thorder Legendre spherical harmonic approximation in sec-tion VA. Finally, the summary, conclusions and outlookare presented in Section VI.

II. SIMULATION DETAILS

We use the Discrete Element Method (DEM) [28] witha simple linear visco-elastic normal contact force lawfn

n = (kδ + γδ)n, where k is the spring stiffness, γnis the contact viscosity parameter and δ or δ are theoverlap or the relative velocity in the normal directionn. The normal force is complemented by a tangentialforce law [28], such that the total force at contact c is:fc = fnn + ftt, where n · t = 0, with tangential forceunit vector t. A summary of the values of the parame-ters used is shown in Table I, with sliding and stickingfriction µ = µsl = µst and rolling– and torsion–torquesinactive (µr = µt = 0). An artificial viscous dissipationforce proportional to the velocity of the particle is addedfor both translational and rotational degrees of freedom,

Value Unit Description

N 9261 [–] Number of particles

〈r〉 1 [mm] Average radius

w 1.5 [–] Polydispersity w = rmax/rmin

ρ 2000 [kg/m3] Particle density

kn 105 [kg/s2] Normal spring stiffness

kt 2.104 [kg/s2] Tangential spring stiffness

µ vary [–] Coefficient of friction

γn 1000 [kg/s] Viscosity – normal direction

γt 200 [kg/s] Viscosity – tangential direction

γbt 100 [kg/s] Background damping – translational

γbr 20 [kg/s] Background damping – rotational

tc 0.64 [µs] Contact duration (average)

TABLE I: Summary and numerical values of particleparameters used in the DEM simulations, where µ, thecontact coefficient of friction is varied in the following.

For more details, see Ref. [28].

resembling the damping due to a background medium,as e.g. a fluid.

A. Simulation set-up and boundary conditions

The simulation set-up is a cuboid volume [16], triaxialbox, with periodic boundaries on all sides. Since careful,well-defined sample preparation is essential, to obtain re-producible results [13], we follow a three-step procedurewhere friction is active in all the preparation stages:(i) Spherical particles are randomly generated in the

3D box with low volume fraction and rather large randomvelocities, such that they have sufficient space and timeto exchange places and to randomize themselves.(ii) This granular gas is then isotropically compressed

to a target volume fraction ν0. The goal is to approacha direction independent, isotropic configuration slightlybelow the jamming volume fraction νc, i.e. the transitionpoint from fluid-like behavior to solid-like behavior [50].(iii) This is followed by a relaxation period at constant

volume fraction to allow the particles to dissipate theirkinetic energy and to achieve a static configuration inmechanical equilibrium, before further preparation or theactual element test is initiated.

B. Isotropic Compression Methods

After the three-step preparation, an isotropic compres-sion test can be initiated to measure isotropic proper-ties and to prepare further initial configurations at dif-ferent volume fractions, with subsequent relaxation, sothat we have a series of different reference isotropic con-figurations, achieved during loading and unloading, asdisplayed in Fig. 1. Note that the initial packings for therespective frictional configurations are inherently differ-

Page 3: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

3

ent since they are prepared with the different friction co-efficients active from the beginning of the first isotropicpreparation stage (stage A in Fig. 1). We only keep ascontrol parameter the volume fraction which is identicalfor the different configurations even though other micro-macro quantities such as pressure and coordination num-ber will be different at a given volume fraction.In the current study, to obtain a homogeneous initial

isotropic configuration, several driving modes have beencompared and these modes are discussed briefly in thefollowing section. Later, for uniaxial tests, unless ex-plicitly mentioned, the wall-driven uniaxial deformationprotocol is applied. We tested the wall-driven againstthe strain-rate driven protocols for some quantities of in-terest to this work and realize that they lead to mostlythe same results – besides some small details (see Sec.II B 5). Note that particular attention must be placed onthe choice of the preparation protocol when other bound-ary conditions or quantities are considered as this conclu-sion may no longer hold. Even though strain-rate drivenproduces more homogeneous systems, we use the wall-driven mode since it more resembles the real experimenttherefore important for future experimental validation ofthis work [9].

1. Wall-driven isotropic compression

In the first method, the periodic walls of the box aresubjected to a strain-controlled motion following a co-sinusoidal law such that the position of e.g., the top wallas function of time t is

z(t) = zf +z0 − zf

2(1 + cos 2πft) (1)

with engineering strain

ǫzz(t) = 1− z(t)

z0, (2)

where z0 is the initial box length and zf is the box lengthat maximum strain, respectively, and f = T−1 is the fre-quency. The maximum deformation is reached after halfa period t = T/2, and the maximum strain-rate appliedduring the deformation is ǫmax

zz = 2πf(z0 − zf)/(2z0) =πf(z0−zf)/z0. The co-sinusoidal law allows for a smoothstart-up and finish of the motion so that shocks and in-ertia effects are reduced. Also, the walls were driven in aquasi-static manner such that the ratio of the kinetic andpotential energy (Ek/Ep ≤ 10−5). By performing slowerdeformations, the energy ratio can be reduced even fur-ther [16].

2. Pressure controlled isotropic deformation

In the pressure controlled mode, the (virtual) periodicwalls of the system are subjected to a co-sinusoidal peri-odic pressure-control until the target pressure is achieved,

for details see [26]. To achieve this, we set the mass ofthe virtual periodic walls of the system mw, to be of theorder of the total mass of the particles in the system,leading to consistent behavior. The pressure controlledmotion of the walls is described by [26]:

mwxw(t) = Fx(t)− pAx(t)− γwx(t), (3)

where Fx(t) is the force due to the bulk material, pAx(t)is the force related to the external load and the last termis a viscous force, which damps the motion of the wall sothat oscillations are reduced. Ax is the area of the wallperpendicular to x where x can be replaced by y or z inEq. 3, for other walls. We find that large values of mw

generally lead to large energy fluctuations/oscillationswhile the final pressure is more rapidly approached forsystems with smaller mw. In contrast, too small mw canlead to violent motions and should be avoided. Addition-ally, we must mention that for our simulations, the sensi-tivity of the system to the wall dissipation is small sincethe simulations are performed in the very slow, quasi-static regime.

3. Homogeneous strain-rate controlled isotropicdeformation

In this method, we apply a homogeneous strain rateto all particles in the ensemble and to the walls in eachtime-step, such that each particle experiences an affine si-multaneous displacement according to the diagonal strainrate tensor:

E = ǫv

−1 0 00 −1 00 0 −1

,

where ǫv (> 0) is the rate amplitude applied until a tar-get maximum volume fraction of e.g., νmax = 0.82 isachieved. The DEM dynamics allows the particles to ap-proach mechanical equilibrium by following the new un-balanced forces that lead to non-affine displacements dueto the new forces at each time-step, or after a relaxationperiod.

4. Swelling of Particles

An alternative isotropic deformation protocol is toallow the particle radii r to slowly ‘grow’ at rate grfrom an initial volume fraction according to the relationdr/dt = grr. The swelling of the particles leads to achange in the volume fraction until the target volumefraction is achieved [30, 36]. During the growth period,the particle mass changes with the radius. Additionally,the volume fraction also changes with time according tothe relation dν/dt = 3νgr, leading to the volume fraction

Page 4: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

4

ν = ν0 exp{3grt} as function of time t. The detailed formof the growth law with time is not relevant here, since allrates are very small.

5. Comparison of driving modes

In summary, comparing the preparation methods, wefind that isotropic quantities like pressure, coordinationnumber or isotropic fabric evolve in a similar fashion forall driving modes. However, the strain-rate controlledisotropic preparation leads to very homogeneous config-urations especially when viewed in terms of the mobilizedfriction. In the wall driven case, we find that friction ismore highly mobilized in the contacts closest to the vir-tual periodic walls of the system leading to slight inho-mogeneities. However, when the particles closest to thewall (up to ≈ 30 % of the box length) are excluded fromthe computation, the resulting probability distributionsas well as the field quantities show negligible differenceswith respect to the data from the full sample analysis.Due to this assessment, we choose here to focus on thewall driven isotropic compression since this more resem-bles experimental set-ups and is especially suitable forthe subsequent uniaxial compression mode. Additionally,the cosinusoidal wall motion allows for a smooth start-upand end of the compression cycle unlike the “kick” (eventhough tiny) to each particle in the strain rate controlledprotocol. To be confident with our conclusions, somedata are checked by comparing them with simulationsperformed with the strain-rate protocol, without comingto different conclusions.

C. Uniaxial Loading and Unloading

After isotropic compression, initial states can be cho-sen from the loading or unloading branch (after relax-ation to allow for kinetic energy dissipation) from whichthe uniaxial test is initiated.As element test, uniaxial compression is achieved by

moving the periodic walls in the z-direction according toa prescribed co-sinusoidal strain path [16], as shown inEq. (1), with diagonal strain-rate tensor

E = ǫu

0 0 00 0 00 0 −1

,

where ǫu is the strain-rate (compression > 0 and decom-pression/tension < 0) amplitude applied in the uniax-

ial mode. The negative sign (convention) of Ezz corre-sponds to a reduction of length, so that tensile deforma-tion is positive. During loading (compression) the vol-ume fraction increases from ν0 (at dimensionless timeτ = t/Tmax = 0) to a maximum νmax = 0.820 (τ = 0.5)

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0 200 400 600 800 1000

ν

Time[ms]

A B C

ν0

νc

νmax

FIG. 1: Evolution of volume fraction as a function oftime. Region A represents the initial isotropic

compression below the jamming volume fraction. Brepresents relaxation of the system to fully dissipate thesystems potential and kinetic energy and C representsthe subsequent isotropic compression up to νmax = 0.820

and then subsequent decompression. Cyan dotsrepresent some of the initial configurations, at differentνi, during the loading cycle; blue stars, for the same νiare different configurations, since obtained during theunloading cycle; both can be chosen for further study.

and reverses back to the original ν0 at the end of the cy-cle (τ = 1), after complete unloading. For more detailson preparation and other parameters, see Ref. [16].

Even though the strain is imposed only on one mo-bile periodic “wall” with normal in the z-direction, whichleads to an increase of compressive stress during com-pression, also the non-mobile x and y directions experi-ence some stress increase as expected for “solid” mate-rials with non-zero Poisson ratio, as discussed in moredetail in the following sections.

However, during decompression the stress on the pas-sive walls is typically smaller than that of the mobile,active wall, as consistent with anisotropic materials andfindings from simulations and laboratory element testsusing the bi-axial tester [23, 55] or the so-called lamb-dameter [24]. One of the main goals of this study is toalso understand the behavior of the packing when com-pression is changed/reversed to tension.

III. DEFINITIONS OF AVERAGEDQUANTITIES

In this section, we present the general definitions ofaveraged microscopic and macroscopic quantities.

Page 5: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

5

A. General Tensor Formulation

To describe and better understand the relationshipsbetween macroscopic quantities, tensors are split up intoisotropic, deviatoric and antisymmetric parts. For a gen-eral decomposition theorem, each tensor can be decom-posed as:

T =1

2(T + T

T ) +1

2(T − T

T ) = T sym + T skew, (4)

where T sym and T skew are the symmetric and antisym-metric parts of the tensor. Since we will focus on the sym-metric part, we further decompose T sym uniquely into itsspherical and deviatoric parts as

T = TvI + TD (5)

with Tv = (1/3)tr(T ) and the traceless deviator TD =T−TvI. The latter contains information about the eigen-system of T , that is identical to the eigensystem of TD

itself.Any (deviatoric) tensor can be transformed using a

transformation matrix R to obtain its diagonal form:

TeigD =

T(1)D 0 0

0 T(2)D 0

0 0 T(3)D

= R

T · TD ·R, (6)

TD = Ti − Tv/3, where Ti’s are eigenvalues of T . Also,

T(1)D , T

(2)D and T

(3)D are the eigenvalues sorted such that,

as convention, T(1)D ≥ T

(2)D ≥ T

(3)D . R = (n1, n2, n3) is

the orthogonal transformation matrix, composed of thecorresponding eigenvectors, which transforms TD to itseigensystem. According to linear algebra, Eq. (6) canalso be expressed as:

TD · nα = TαDnα (7)

with TαD and nα the α-eigenvalue and eigenvector of TD,

respectively. The symbol “·” represents the inner productof the tensor TD and the vector nα which leads to avector parallel to nα.In the following, we provide a consistent decomposition

for strain, stress and fabric tensors. We choose here todescribe each tensor in terms of its isotropic part (firstinvariant) and the second (J2) and third (J3) invariantof the deviator:

J2 =1

2

[

(T(1)D )2 + (T

(2)D )2 + (T

(3)D )2

]

(8)

J3 = det(TD) = T(1)D T

(2)D T

(3)D (9)

J3 can further be decomposed as J3 =

T(1)D T

(2)D (−T (1)

D − T(2)D ), since we are dealing with

deviators.

B. Strain

For any deformation, the isotropic part of the infinites-imal strain tensor ǫv (in contrast to the true strain εv) isdefined as:

ǫv = ǫvdt =ǫxx + ǫyy + ǫzz

3=

1

3tr(E) =

1

3tr(E)dt,

(10)where ǫαα= ǫααdt with αα = xx, yy and zz as the di-agonal elements of the strain tensor E in the Cartesianx, y, z reference system. The integral of 3ǫv denoted

by εv = 3∫ V

V0

ǫv, is the true or logarithmic strain, i.e.,

the volume change of the system, relative to the initialreference volume, V0 [14].Several definitions are available in literature [48] to de-

fine the deviatoric magnitude of the strain. Here, we usethe objective definition of the deviatoric strain in termsof its eigenvalues ǫd

(1), ǫd(2) and ǫd

(3) which is indepen-dent of the sign convention.The deviatoric strain is defined as:

ǫdev =

(

ǫ(1)d − ǫ

(2)d

)2

+(

ǫ(2)d − ǫ

(3)d

)2

+(

ǫ(3)d − ǫ

(1)d

)2

2,

(11)where ǫdev ≥ 0 is the magnitude of the deviatoric strain.Note that the wall motion is strain controlled and the

infinitesimal strain corresponds to the external appliedstrain. Hence the eigenvalues for the strain tensor are inthe Cartesian coordinate system (thus no transformationis needed). For the purely isotropic strain, ǫISO = ǫvI,with ǫdev = 0, which is direction independent by defini-tion. The corresponding shape factor for strain Λ(−ǫ), is

represented as the ratio Λ(−ǫ) := ǫ(2)d /ǫ

(1)d .

C. Stress

From the simulations, one can determine the stress ten-sor (compressive stress is positive as convention) compo-nents:

σαβ =1

V

p∈V

mpvpαvpβ −

c∈V

f cαl

, (12)

with particle p, mass mp, velocity vp, contact c, forcef c and branch vector lc, while Greek letters representcomponents x, y, and z [16, 27]. The first sum is thekinetic energy density tensor while the second involvesthe contact-force dyadic product with the branch vector.Averaging, smoothing or coarse graining [52, 53] in thevicinity of the averaging volume, V , weighted accordingto the vicinity is not applied in this study, since aver-ages are taken over the total volume. Furthermore, sincethe data in this study are quasi-static, the first sum can

Page 6: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

6

mostly be neglected. The isotropic stress is denoted ashydrostatic pressure:

p = σv =1

3tr(σ) (13)

As already mentioned, we will focus on the eigenvaluesof the deviatoric stress tensor λsi = σD

i = σi−p, as definedin section III A, with the principal directions being thesame for σ and σ

D. The (scalar) deviatoric stress for our3D uniaxial simulations is:

σdev =

(λs1 − λs2)2 + (λs1 − λs3)

2 + (λs2 − λs3)2

2, (14)

The deviatoric stress ratio, sdev = σdev/p, quantifiesthe “stress anisotropy” - where σdev =

3Jσ2 , with Jσ

2

the second invariant of the deviatoric stress tensor. Thethird stress invariant Jσ

3 = λs1λs2λ

s3 = λs1λ

s2(−λs1 − λs2) =

λs13(−Λ1

σ − (Λ1σ)2) can be replaced by the shape fac-

tor Λσ := λs2/λs1, which switches from +1 at maximum

uniaxial loading to -1/2 after some unloading as will beshown below.

D. Structural (Fabric) Anisotropy

Besides the stress of a static packing of powders andgrains, an important microscopic quantity of interest isthe fabric/structure tensor. For disordered media, theconcept of a fabric tensor naturally occurs when the sys-tem consists of an elastic network or a packing of discreteparticles. A possible expression for the components of thefabric tensor is provided in [27, 31]:

F ναβ = 〈F p〉 = 1

V

p∈V

V p

N∑

c=1

ncαn

cβ , (15)

where V p is the particle volume of particle p which liesinside the averaging volume V , and nc is the normal vec-tor pointing from the center of particle p to contact c.F ναβ are thus the components of a symmetric rank two

3 × 3 tensor. In a large volume with some distributionof particle radii, the relationship between the trace offabric, volume fraction ν and the average coordinationnumber C is given by 3Fv

ν := F ναα = g3νC, as first re-

ported in [29] and also confirmed from our wider friction(µ) data. The term g3 corrects for the fact that the co-ordination number for different sized particles is propor-tional to their surface area such that for a monodispersepacking g3 = 1 and for a polydisperse packing g3 > 1[14, 31, 41].A different formulation for the fabric tensor considers

simply the orientation of contacts normalized with thetotal number of contacts Nc, as follows [25, 35, 39]:

Fαβ =1

Nc

N∑

c=1

ncαn

cβ , (16)

The relationship between Eq. (15) and Eq. (16) is:

Fαβ =F ναβ

g3νC=

3F ναβ

Fv. (17)

We can define the deviatoric tensor FD and calculatethe eigenvalues λfi = Fi − Fv/3 with Fv = 1, and Fi theeigenvalues of the deviatoric fabric based on Eq. (16).We assume that the structural anisotropy in the system

is quantified (completely) by the anisotropy of fabric, i.e.,the deviatoric fabric, with scalar magnitude similar toEqs. (11) and (14) as:

Fdev =

(λf1 − λf2 )2 + (λf1 − λf3 )

2 + (λf2 − λf3 )2

2, (18)

proportional to the second invariant of FD, Fdev =

3JF2 , where λf1 , λ

f2 and λf3 are the three eigenvalues

of the deviatoric fabric tensor.Alternatively, a simpler definition of the deviatoric fab-

ric for an axial symmetric element test takes into accountthe difference between the fabric eigenvalue of the maincompressive (axial) direction and the average values inthe isotropic plane as follows:

F ∗

dev = λf1 − λf2 + λf32

. (19)

Note that if λf2 = λf3 , Eqs. (18) and (19) coincide.Analogous to the definitions in Eqs. (18) and (19), Fdev

and F ∗

dev can also be defined using the definition of thefabric presented in Eq. (15).

E. Eigenvector Orientation

Due to the axial symmetry of the uniaxial compressionmode, the orientation of the eigenvectors of stress andfabric can be defined with reference to the main com-pressive z-direction as:

θα = arccos(

n(α) · z

)

(20)

where z is the unit vector in the z-direction. Addition-ally, orientations are projected such that they lie withinthe range to π/2.

IV. RESULTS AND OBSERVATIONS

In this section, as results of the current study, firstwe will discuss the influence of friction on the evolutionof stress and structural anisotropy as functions of de-viatoric strain during loading and unloading. To com-plement these results, we investigate the magnitude and

Page 7: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

7

orientation of the eigenvalues of stress and fabric duringloading and unloading and their respective shape factors.To gain an insight into the relationship between the nor-mal and tangential force on the macroscopic stress andstructure, we report briefly their probability density func-tions (pdfs) for different frictional systems, as well as theforce intensity weighted by the contact state. Finally, wepresent a 6th order harmonic approximation of the po-lar representation of contacts and forces to describe theaxial-symmetric structural anisotropy, relating fabric tothe pdfs.Isotropic quantities during loading and unloading for

various deformation paths were presented in Ref. [16] forfrictionless particles and will not be detailed here. Weonly note that the coordination number and the hydro-static pressure have been shown to scale quantitativelydifferently with isotropic strain but behave in a very sim-ilar fashion irrespective of the deformation mode applied.The pressure has also been shown to be coupled to thedeviatoric strain via the structural anisotropy. The ef-fects of polydispersity on the evolution of the isotropicquantities have also been extensively studied in Ref. [20]for various deformation paths. The isotropic quantities,namely pressure, coordination number and fraction ofrattlers show a systematic dependence on the deforma-tion mode and polydispersity via the respective jammingvolume fractions.In the current work, we will focus on the non-isotropic

quantities and their evolution with respect to the devia-toric strain.

A. Deviatoric Stress and Fabric

Under uniaxial compression, not only does shear stressbuild up, but also the anisotropy of the contact andforce networks develops, as it relates to the creation anddestruction of new contacts [16]. We term the devia-toric part of the stress tensor and its microscopic force-direction dependence as the “stress anisotropy”, in paral-lel to the contact direction-dependency of the structuralanisotropy.The deviatoric stress ratio, sdev = σdev/p is shown in

Figs. 2(a) and 2(b) for a frictionless (µ = 0) and sev-eral frictional (µ = 0.01, 0.02, 0.05, 0.1, 0.2, 0.3, 0.5and 1.0) systems during uniaxial loading and unloading,respectively. As the deviatoric strain applied to the sys-tem is increased during uniaxial loading, the deviatoricstress ratio initially grows for all the friction coefficientsshown. In some cases (for small µ), the maximal sdevis reached before the maximum deviatoric strain applied(εmax

dev =0.1549) is reached. For some of the configura-tions studied, an asymptote (or steady state) is observedin which further application of deviatoric strain does notlead to visible further increase/decrease in the deviatoricstress. At the maximum applied deviatoric strain, weobserve that not all configurations (especially the high-est friction coefficients) have reached full saturation. For

the systems with lower microscopic friction coefficients,a slight decrease of the deviatoric stress ratio for largerdeviatoric strains is seen. The slope of the deviatoricstress ratio, which represents its growth rate shows a de-creasing trend with increasing friction. Recall that theinitial packings are different since they are prepared withdifferent friction coefficients. Due to this, the pressureincreases with increasing friction while the coordinationnumber decreases with friction. The slope of the devia-toric stress ratio in Fig. 2(a), related to the initial shearstiffness of the isotropic packing is proportional to thesetwo quantities [12, 32, 51].

The evolution of the deviatoric stress during unload-ing (after strain reversal) is presented in Fig. 2(b). Notethat due to the square-root definition used in Eq. (14),the deviatoric stress remains positive [57]. During devi-atoric unloading, sdev begins to decrease until the sys-tem approaches an isotropic stress configuration, wheresdev = 0. The εdev values where sdev ≈ 0 consistentlydecrease with increasing friction – as consistent with thetrend of the maximum sdev values reached during uniax-ial loading at larger εdev for stronger friction. For systemswith large friction coefficients (µ = 0.3, 0.5 and 1.0), theεdev values at which sdev = 0 are closer to each otherthan for weakly frictional systems – see Fig. 8 below.

Along with the deviatoric stress ratio, for a character-ization of the contact network of the particles, we plotthe deviatoric fabric magnitudes Fdev of the systems dis-cussed above as function of the deviatoric strain duringuniaxial loading and unloading in Figs. 2(c) and 2(d), re-spectively. In Fig. 2(c), the deviatoric fabric magnitudebuilds up from different (random, but small) initial val-ues and reaches different maxima within the same rangeof deviatoric strain (εdev ≈ 4−6%). For larger strains, weobserve a decrease in the structural anisotropy towardszero. Interestingly, for systems with higher friction coef-ficients (µ = 0.3, 0.5 and 1.0), after the decrease in thestructural anisotropy, further loading in the axial direc-tion leads to a (small) second increase of the deviatoricfabric, until at maximum compression, the deviatoric fab-ric again reaches a local maximum. This indicates thatmore contacts are created in the axial compressive direc-tion compared to the horizontal plane at the beginningof the loading cycle. At the first maximum (εdev ≈ 0.06),the material behavior changes such that the number ofcontacts created in the horizontal plane becomes higherwith respect to the vertical plane. This trend reversesagain as maximum compression is reached for systemswith higher friction coefficients. This interesting behav-ior will be further discussed when we analyze the mag-nitude and orientation of the respective eigenvectors inSection IVB.

After strain reversal, as presented in Fig. 2(d), the ini-tial isotropic state is not recovered – a clear signature ofhistory dependence and structural anisotropy being in-dependent of (or decoupled from) the deviatoric stressratio. Additionally, a strong difference can be seen in thefabric response of systems with lower and higher friction,

Page 8: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

8

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

s dev

εdev

µ=0.0

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(a)

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0

s dev

εdev

µ=0.0

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(b)

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

Fde

v

εdev

µ=0.0

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(c)

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0

Fde

v

εdev

µ=0.0

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(d)

FIG. 2: The deviatoric stress ratio plotted as function of deviatoric strain during uniaxial (a) loading and (b)unloading. The corresponding plots of the deviatoric fabric plotted during uniaxial (c) loading and (d) unloading,

for different microscopic friction coefficients.

respectively. As we will see later, the orientation of theeigenvalues of these systems provide interesting insightsinto these observations.

In general, comparing the evolution of deviatoric stressratio and deviatoric fabric, we observe a strongly decou-pled and non-linear qualitative behavior with the linearcontact model used in this study. This confirms that thenon-linearity observed is a peculiarity of the deformationmode and the structure of the packing.

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6

µ

sdevmax Fdev

max

FIG. 3: Trend of the peak deviatoric stress and peakdeviatoric fabric with increasing microscopic frictioncoefficient µ under uniaxial loading, given a maximalstrain εmax

dev = 0.1549. smaxdev values for µ > 0.1 are taken

at εmaxdev since no clear maximum is achieved. Dashed

line indicates µmacro = µ.

In Fig. 3, we plot the maximum deviatoric stress ratioand maximum deviatoric fabric reached from Figs. 2(a)and 2(c) for the respective friction coefficients. Inter-estingly, the maximum deviatoric stress ratios increasewith increasing friction coefficient until µ ≈ 0.25, whereit peaks at smax

dev ≈ 0.43 and subsequently decrease forhigher friction coefficients. This maximum reached istermed the “macroscopic friction coefficient”, µmacro :=smaxdev [16], representing the macroscopic mobilized fric-tion, i.e. shear resistance of the material. We note thatthe maxima reached are higher than the microscopic fric-tion coefficient for systems with low friction, between µ= 0 and 0.4, while for higher friction, the maxima arelower [49].In Fig. 3, we also show the trend of the maximum

structural anisotropy reached, Fmaxdev , with increasing fric-

tion. Besides the increase between, µ = 0 and 0.01,the maximum deviatoric fabric shows a decreasing trendwith increasing friction and saturates at Fmax

dev ≈0.025for the highest friction coefficients. In comparison, thestructural anisotropy is much smaller than the deviatoricstress ratio and it is reached at comparable εdev, whereasthe extreme stress anisotropy is reached much later forhigher µ.

B. Eigenvalues and Eigenvectors of Stress andFabric

In this section, we will discuss the magnitude of theeigenvalues of deviatoric stress and deviatoric fabric dur-

Page 9: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

9

-1000

-500

0

500

1000

1500

2000

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

λs

εdev

123

(a)

-1000

-500

0

500

1000

1500

2000

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0

λs

εdev

123

(b)

0

20

40

60

80

100

120

140

160

180

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

θs

εdev

123

(c)

0

20

40

60

80

100

120

140

160

180

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0

θs

εdev

123

(d)

FIG. 4: Eigenvalues of stress for µ = 0.1 plotted as functions of the deviatoric strain for (a) loading and (b)unloading along with their corresponding orientations with respect to the compressive direction during uniaxial (c)

loading and (d) unloading.

-0.06

-0.04

-0.02

0

0.02

0.04

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

λf

εdev

123

(a)

-0.06

-0.04

-0.02

0

0.02

0.04

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0

λf

εdev

123

(b)

0

20

40

60

80

100

120

140

160

180

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

θf

εdev

123

(c)

0

20

40

60

80

100

120

140

160

180

0.14 0.12 0.10 0.08 0.06 0.04 0.02 0

θf

εdev

123

(d)

FIG. 5: Eigenvalues of the deviatoric fabric for µ = 0.1 plotted as functions of the deviatoric strain for (a) loadingand (b) unloading along with their corresponding orientations with respect to the compressive direction during

uniaxial (c) loading and (d) unloading.

ing uniaxial loading and unloading as well as the orienta-tion of the eigenvectors. As reference and representativeexample, we will show the data for only one of the co-efficients of friction (µ = 0.1) and discuss in words the

interesting trends for the others. Finally, we will couplethe observations to the evolution of stress and structuralanisotropies presented in section IVA.

In Figs. 4(a) and 4(b), we plot the eigenvalues of the

Page 10: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

10

deviatoric stress for the frictional system with µ = 0.1during loading and unloading against deviatoric strainεdev. During loading λs1 which corresponds to the stresseigenvalue of the axial compression direction increaseslinearly from 0 and remains positive while the eigenval-ues λs2 and λs3 of the two non-mobile direction are neg-ative and very similar in magnitude. During unloading,λs1 decreases but remains positive; at εdev ≈ 0.075, alleigenvalues become zero and then switch order, so thatthe axial direction eigenvalue is becoming increasinglynegative. The intermediate λs2 then gets identical to λs1,both growing to positive values. The orientation of thecorresponding eigenvectors during loading and unload-ing are shown in Figs. 4(c) and 4(d). At εdev = 0, theorientations are different and random which is an indi-cation of the almost isotropic initial configuration. Withincreasing strain, θs1, which corresponds to the orienta-tion of the compressive stress eigenvalue, converges toθs= 0◦ and remains until the end of the loading path.During this period, the stress and strain eigenvectors aresaid to be colinear with respect to each other. On theother hand, the orientation θs2 and θs3 of the other eigen-values also drops to θs= 90◦ showing a perpendicularalignment with respect to the compression direction. Af-ter strain reversal, the eigendirections of stress do notinstantaneously respond to the directional change untilat εdev ≈ 0.10 where θs1 begins to increase to 90◦ andfinally reaches at εdev ≈ 0.03. Accordingly, θs3 drops to0◦, while θs2 remains close to 90◦ all the time.The corresponding eigenvalue and eigenvector orienta-

tions of the deviatoric fabric for µ = 0.1 are presentedin Figs. 5(a) and 5(b) during uniaxial loading and un-loading. Similar to the eigenvalues of stress, the major

eigenvalue λf1 , remains positive while the two lower eigen-

values are negative. In contrast to stress, λf1 increasesand reaches a peak at εdev ≈ 0.05 after which it beginsto decrease towards zero as the maximum strain is ap-

proached. Also, λf2 and λf3 are not identical, i.e., λf3 has a

slightly higher magnitude than λf2 . This is an indicationof the existence of anisotropy in the plane perpendicu-

lar to λf1 even though the stress picture shows isotropy.At maximum deviatoric strain, however, the magnitudesof all the eigenvalues are close to zero. After strain re-

versal, λf1 and λf2 show an increasingly positive trendfrom εdev ≈ 0.08 but are not exactly identical in magni-

tude while λf1 is negative and consistently decreases fromεdev ≈ 0.08 until the end of the decompression cycle.Similar to the stress, the orientations of the fabric com-

ponents are interesting. Starting from random values, θf1decreases and is close but distinct from 0◦ during load-

ing, while θf2 and θf3 are close to 90◦ during the same pe-

riod. This indicates that θf1 is not fully aligned with thestrain eigenvector with the deviation showing the non-colinearity. After strain reversal, a delay can be seen be-

fore θf1 and θf3 transit to 90◦ and 0◦, respectively, while

θf2 remains close to 90◦.Additionally, to fully describe the tensors, one can cal-

Shape factor τ ≈ 0 τ ≈ 0.5 τ ≈ 1Λσ = λs

2/λs1 Random -1/2 1

Λf = λf2/λ

f1 Random -1/2 1

Λ(−ǫ) = ǫ(2)d /ǫ

(1)d Undefined -1/2 1

TABLE II: Shape-factors of stress and fabric in therespective tensor eigensystem at the beginning,maximum and end of uniaxial compression.

culate the respective shape factors for stress and fabric,respectively, as the ratio of the eigenvalues as shown inTable II at the initial, maximum and end of the uniaxialcompression–decompression cycle.

In the following analysis, we will investigate how theorientation changes with increasing the microscopic fric-tion coefficient and the relationships with the force net-work.

In Figs. 6(a) and 6(b), we plot the orientations of the

first eigenvectors of stress θs1 and fabric θf1 for all contactsand different friction coefficients, respectively. The initialvalue of θs1 is random at the beginning of the loading pathfor the different friction coefficients. As loading begins,θs1 decreases and at εdev ≈ 0.02, θs1 ≈ 0◦ for all friction.The relaxation rate (data scaled with the initial value ofthe respective θs1), shown as an inset on a log-scale is non-systematic for the different friction coefficients possiblydue to the initial isotropic configuration. Note that sincethe angle θs1 does not exactly decrease to zero since θs1is always positive even though it fluctuates around zero.Observing the behavior of the eigenvectors nx and ny ofthe largest eigenvalue, we find that during loading, theyapproach zero (aligned with the compression direction)and remain until maximum compression. After strain re-versal, a slight delay is seen before the vectors finally flipback to the plane [15]. After strain reversal at εdev =0.16, the response of θ1(s) is slow and it only begins toincrease at εdev ≈ 0.12 for µ = 0. It is interesting tonote that the delay time increases with friction and pos-sibly due to the higher maximum deviatoric stress val-ues reported with increasing friction. The correspondingorientation of the major eigenvector of fabric θ1(f) forall contacts and different friction coefficients also startsfrom different random values before decreasing to 0◦ withincreasing loading. Surprisingly at εdev = 0.08, for theconfigurations with lower friction (µ =0, 0.01, 0.02 and0.05), θ1(f) remains close to 0◦ while those with higherfriction (µ =0.2, 0.3, 0.5 and 1.0) begin to increase to-wards 90◦ as we approach maximum compression. Thisindicates that the orientations and build-up of contactsfor systems with lower/higher friction behave in oppositefashion to each other and makes clear the reason for thedecrease seen in the deviatoric fabric evolution in Fig.5(a). At the lower friction case, with increasing loading,contacts are mostly built along the main compression di-rection. However with increasing friction, a ‘saturation’of contact build-up in the vertical direction sets in andan increasing number of contacts begin to build-up in the

Page 11: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

11

horizontal direction. As strain is reversed, the eigenvec-tor orientation for systems with low friction increases to90◦ while a decrease before an increase to 90◦ is seen forsystems with higher friction.

To further understand this interesting observation wesub-divide the respective systems into strong and weakcontacts and we plot the orientation of the stress andfabric eigenvector corresponding to the compression di-rection for the two sub-divisions. Strong contacts aretermed as those whose normal force intensity is greaterthan the mean normal force while those with lower in-tensity with respect to the mean normal force are termedweak.

We plot the orientation of the major direction eigen-vector of stress and fabric respectively in Figs. 6(c) and6(d) for strong contacts. From Fig. 6(c), the orientationof the strong contact main eigenvector of stress and fab-ric behaves in a similar fashion as the total contact inthe ensemble. This is consistent with earlier findings [47]where the strong contacts have been observed to carrymost of the load during deformation. Interestingly andin contrast to the observation for all contacts, the fabriceigenvalue for systems with both low and high friction allstay close to 0◦ during loading and initial unloading.

Next, the orientation of the main eigenvector of stressand fabric for weak contacts is shown in Figs. 6(e) and6(f). Similar to the strong contacts, the stress and fabricorientation of weak contacts behave in a similar fashionbut in contrast are mostly oriented at 90◦ during loading.During unloading, the orientation tends towards 0◦.

Comparing Figs. 6(b), 6(d) and 6(f), it can be seenthat strong contacts predominate for the system withvery low friction while for higher friction, the orientationof the weak contacts play a much significant role.

We also plot the respective shape factors as ratio ofthe eigenvalues of stress and fabric for some exemplaryfriction coefficients during uniaxial loading and unload-ing in Fig. 7. For stress, shown in Fig. 7(a), beginningfrom random values, Λσ decreases to -1/2 during loadingand reverses to 1 at the end of the unloading cycle. Therates of change during loading and unloading are almostidentical, for different µ while during unloading, the devi-atoric strain at which the increase occurs decreases withincreasing friction. As with the stress, the shape factorof fabric Λf , shown in Fig. 7(b), also begins from randomvalues and during loading approaches Λf ≈ −1/2 withstronger fluctuations for higher friction coefficients. Atthe end of unloading however Λf approach unity.

In Fig. 8, we plot the deviatoric strains at which themajor eigenvalues θs1 cross 45◦ during unloading for dif-ferent friction coefficients. Additionally, we also plot thedeviatoric strains at which the deviatoric stress ratio,deviatoric fabric and the stress shape factor cross zerofrom Figs. 2(b), 2(d) and 7(a), respectively. As shown,the transition point decreases non-linearly with increas-ing friction. All data originating from the stress tensor,namely the major eigenvalue of stress, its orientation andthe stress shape factor all collapse on each other. On the

other hand, it is not surprising that the transition pointsfor the fabric quantities are slightly off since the fabricbehaves differently from the stress. The definition of thefabric tensor takes into account only the normal direc-tions and does not include the strong tangential contribu-tions to the contacts. Therefore, as friction is increased,the deviations can be stronger.In the following section, we will investigate in more

detail the fraction of weak and strong contacts in thesesystems and discuss their interplay and relation to theobservations on the orientations of the strong and weakcontacts. For clarity and to better view the evolution ofthe quantities, instead of the deviatoric strain ǫdev, wewill study the evolution of the quantities against dimen-sionless time τ = t/T – where T is the simulation time.

C. Friction Mobilization

Mobilization of contact friction, during uniaxial defor-mation of the bulk material, is quantified by the factorft/µfn ≤ 1 for each contact. The tangential forces growtowards their limit and support larger shear stress; fortangential forces at/above the Coulomb limit, i.e., at fullymobilized friction, sliding sets in and rearrangements ofcontacts can lead to new, more stable configurations. Ithas been observed [44] that sliding is mostly active atweak contacts (termed weak sliding, wsl), while strongercontacts stay in the sticking regime and sustain largerfriction forces while being less mobilized (termed strongsticking sst). We refer to this as the ws–rule. Weak andstrong contacts are defined relative to the average normalforce at each timestep;

f∗ = fn/〈fn〉 < 1 (21)

are termed weak and

f∗ = fn/〈fn〉 > 1 (22)

are termed strong [44], with dominating sliding and stick-ing, respectively.As we will see shortly, we find that this friction mo-

bilization rule may not strictly hold in certain cases, asthere may be a considerable number of weak contactswith friction not fully mobilized (termed weak sticking,wst), as well as strong contacts fully mobilized (termedstrong sliding, sst).As representative examples, in Fig. 9, we track two

different contacting pairs during uniaxial loading and un-loading of the system with µ = 0.1 and study the forceintensity and friction mobilization as they evolve as func-tion of the dimensionless time τ . For the first contactpair shown in Fig. 9(a), during the first stages of load-ing, the contact is weak since f∗ < 1; friction is fullymobilized and sliding occurs at the contact, i.e. weakcontacts tend to full friction mobilization. For a short pe-riod at τ ≈ 0.2, the contact becomes stronger and ft/µfn

Page 12: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

12

0 10 20 30 40 50 60 70 80 90

100

0 0.04 0.08 0.12 0.16

θ 1s

εdev

10-3

10-2

10-1

1

10

0 0.02 0.04

0.12 0.08 0.04 0

εdev

µ=0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.3µ=0.5µ=1.0

(a)

0 10 20 30 40 50 60 70 80 90

100

0 0.04 0.08 0.12 0.16

θ 1f

εdev

0.12 0.08 0.04 0

εdev

µ=0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.3µ=0.5µ=1.0

(b)

0 10 20 30 40 50 60 70 80 90

100

0 0.04 0.08 0.12 0.16

θ 1s(

stro

ng)

εdev

10-3

10-2

10-1

1

10

0 0.02 0.04

0.12 0.08 0.04 0

εdev

µ=0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.3µ=0.5µ=1.0

(c)

0 10 20 30 40 50 60 70 80 90

100

0 0.04 0.08 0.12 0.16

θ 1f(

stro

ng)

εdev

0.12 0.08 0.04 0

εdev

µ=0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.3µ=0.5µ=1.0

(d)

20

30

40

50

60

70

80

90

100

0 0.04 0.08 0.12 0.16

θ 1s(

wea

k)

εdev

0.12 0.08 0.04 0

εdev

µ=0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.3µ=0.5µ=1.0

(e)

30

40

50

60

70

80

90

100

0 0.04 0.08 0.12 0.16

θ 1f(

wea

k)εdev

0.12 0.08 0.04 0

εdev

µ=0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.3µ=0.5µ=1.0

(f)

FIG. 6: Orientation of the largest positive (a) Stress eigenvector for all contacts (b) fabric eigenvector for allcontacts (c) Stress eigenvector for strong contacts (d) Fabric eigenvector for strong contacts (e) Stress eigenvector

for weak contacts (f) Fabric eigenvector for weak contacts plotted against dimensionless time for different coefficientof friction.

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

0 0.04 0.08 0.12 0.16

Λσ

εdev

0.12 0.08 0.04 0

εdev

µ=0

µ=0.05

µ=0.2

µ=0.5

µ=1.0

(a)

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1

0 0.04 0.08 0.12 0.16

Λf

εdev

0.12 0.08 0.04 0

εdev

µ=0µ=0.05

µ=0.2µ=0.5µ=1.0

(b)

FIG. 7: Shape factors of (a) stress and (b) fabric asfunction of the deviatoric strain for some exemplary

friction coefficients.

correspondingly reduces (with strong fluctuations) indi-cating a strong contact where sticking predominates. Atτ ≈ 0.36, the contact between this particle pair is lost(opened) and is only recovered at τ ≈ 0.7, where it can

0

0.02

0.04

0.06

0.08

0.1

0.12

0 0.2 0.4 0.6 0.8 1

ε dev

µ

θ1s= 45°

sdev= 0Fdev= 0

Λσ= 0θ1

f(strong)= 45°

(a)

FIG. 8: Strains at which the orientations of the stresseigenvectors cross θ = 45◦ and at which the deviatoricstress ratio, deviatoric fabric and the stress shape factorcross zero for frictions µ =0, 0.01, 0.02, 0.05, 0.1, 0.2,

0.5 and 1.0.

again be classified as weak sliding (wsl) contact. As theend of the compression cycle is reached, the contact in-tensity increases and ft/µfn decreases, with strong fluc-tuations again, and sometimes sliding. In general, thews–rule is mostly true for this contact pair except duringthe transition from weak to strong where some fluctua-tions in ft/µfn can be seen, transitions from sliding tosticking can happen for weak contacts (wst) well below

Page 13: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

13

f∗ = 1 during increase of f∗, but also sliding can happenfor strong contacts (ssl).

The second contact pair shown in Fig. 9(b) is evenmore interesting. Like the first particle pair, the secondpair also begin as a weak sliding contact and f∗ growsuntil τ ≈ 0.15, where it becomes strong. Interestingly,while the contact remains very strong for almost all ofthe loading–unloading cycle, friction is highly mobilizedft/µfn remains close to 1.

Since studying just two contact pairs within an en-semble containing tens of thousands of contacts providesvery little information, we first extract the total fractionof weak and strong contacts in the system. In Fig. 10(a),we plot the total proportion of weak contacts with ref-erence to the total number of contacts for the differentfriction coefficients (which was studied in detail in Refs.[16, 20] so that those data are not shown here). Surpris-ingly, as with the orientation of the largest eigenvalueof fabric for weak and strong forces plotted in Fig. 6,we see a clear difference between the fraction of weakand strong contacts. In the following, we will discuss indetail the observations for weak contacts – which haveopposite trends as the observations for strong contacts.

The first observation from Fig. 10(a) is that a greaterfraction (over 50%) of the contacts in the respective sys-tems are weak – an indication that fewer contacts carry alarger than average proportion of the load in the system,which is due to the shape of the force probability den-sity function P (f∗), see Section IVD. Secondly, for sys-tems with lower friction, the fraction of weak contacts atthe beginning of the loading cycle is significantly higherthan for higher friction, meaning that the load is moreevenly (not exactly proportionally) distributed betweenweak and strong contacts for systems with higher fric-tion coefficient. With increasing loading, while the totalnumber Ctot strongly increases (not shown), the fractionof weak contacts decreases for packings with lower fric-tion coefficients, and increases for those with higher fric-tion. Also, the decrease of weak contacts with increasingloading for lower friction systems is stronger and occursearlier than the increase for systems with higher friction.At maximum loading τ = 0.5, the proportion of weakcontacts are close for all friction coefficients with slightlyhigher fraction for the highest friction coefficients µ =0.5 and 1.0. This observation, that the packings withhigher friction behave in a qualitatively different fash-ion, is consistent with the earlier observation in Fig.6(b),where the difference in orientation of strong/weak con-tacts for low/high friction coefficients can be seen too.

It is surprising that the fractions of weak contacts areclose for systems with lower friction and evolve in a sim-ilar (almost symmetric) fashion during loading and un-loading. For µ = 0.01 and 0.05, the fractions of weakcontacts at the end of unloading are slightly lower thanat the beginning of loading. With increasing friction,the fractions of weak contacts at the end of unloadingare higher than at the beginning of loading; the anti-symmetry between the loading and unloading phases is

more visible for µ ≥ 0.1.

To evaluate the proportion of weak and strong con-tacts contributing to sliding and sticking at contacts, weplot in Fig. 11, the number of weak sliding (

wsl) andstrong sticking (

sst) contacts with respect to the thetotal weak (

w) and strong (∑

s) contacts, respectively.From Fig. 11(a), the fraction of weak sliding contactsgrows during loading and reaches a peak before it beginsto decrease towards zero as maximum loading (τ = 0.5)is approached. The initial growth rate of the weak slidingcontacts and the peak reached decreases with increasingfriction but all approach zero at τ = 0.5 because thedeformation rate decreases to zero before reversal. Dur-ing unloading, a second growth phase of the weak slid-ing contacts is seen and the maximum reached is higherthan that reached during loading – thus leading to a non-symmetry around τ = 0.5. Additionally, only a smallproportion (much less than ≈ 50%) of the total weakcontacts are sliding. This indicates that even though anincrease in the number of weak sliding contacts is seenduring loading and unloading, more and more weak con-tacts stick (ft/µfn < 1) for increasing µ.

In contrast to the weak sliding contacts, the fractionof strong sticking contacts, as presented in Fig. 11(b) de-creases during loading until it reaches a minimum beforean increase towards τ = 0.5 can be seen. The rate of de-crease and the minima reached decrease with increasingfriction and the minima are lower during unloading, i.e.all data are non-symmetric around τ = 0.5.

In Fig. 12, we plot the number of weak sliding (∑

wsl)and strong sticking (

sst) contacts with respect to thetotal sliding and sticking contacts, respectively. In Fig.12(a), we confirm that a higher proportion (> 0.5) of thesliding contacts are weak [5, 6, 44, 45]. The proportionsof weak sliding contacts for µ = 0.01 − 0.3 are almostidentical and decrease during loading. During unload-ing, however, the proportions of weak sliding contactsbehave differently with increasing friction. We again ob-serve the non-symmetry of the loading and unloadingdata. In Fig. 12(b), we plot the fraction of strong stick-ing contacts with respect to the total sticking contacts.A little less than 50 % of the sticking contacts are strong.The fractions of strong sticking contacts increase initiallyduring loading and later decreases as maximum compres-sion is approached. The fraction of strong sticking con-tacts show a decreasing trend at τ = 0.5 with increas-ing µ. During unloading, the fractions of strong stickingcontacts increase and later decrease towards the end ofthe unloading branch. With increasing friction, the non-symmetry of the data decreases. For the highest frictioncoefficients, the fraction of strong sticking contacts dur-ing loading is slightly more than those present duringunloading.

In summary, strong and weak forces have been ana-lyzed along with the level of friction mobilization. It hasbeen shown that a higher proportion of the total contactsin the system are weak, irrespective of the friction coef-ficient. Among these weak contacts, the contacts which

Page 14: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

14

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

τ

fn/<fn> ft/µfn

(a)

0

0.5

1

1.5

2

2.5

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

τ

fn/<fn> ft/µfn

(b)

FIG. 9: Tracking ft/µfn and f∗ = fn/〈fn〉 for two single particle pairs randomly selected from the ensemble duringcompression and decompression where τ is the dimensionless time. (a) Particle pair 1 (b) Particle pair 2.

0.525

0.53

0.535

0.54

0.545

0.55

0.555

0.56

0.565

0.57

0.575

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Cw

/Cto

t

τ

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(a)

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Csl

/Cto

t

τ

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(b)

FIG. 10: Proportion of (a) weak contacts (c) sliding contacts with respect to the total number of contacts duringuniaxial loading- and unloading-cycle for different friction coefficients.

0

0.1

0.2

0.3

0.4

0.5

0.6

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Σwsl

/Σw

τ

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(a)

0.6

0.65

0.7

0.75

0.8

0.85

0.9

0.95

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Σsst

/Σs

τ

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(b)

FIG. 11: Fraction of (a) weak sliding contacts (wsl) and (b) strong sticking contacts (sst) with respect to the totalnumber of weak (

w) and strong (∑

s) contacts, respectively, during uniaxial loading- and unloading for differentfriction coefficients.

are sliding are less in number compared to the stickingcontacts. In contrast, when the total sliding contacts areconsidered, a higher proportion of them are weak, as alsoreported in earlier literature.

D. Probability density function

To better understand the relationship between contactforces and the macroscopic stress and structure, we firststudy the probability density function of normal contactforces in different directions.

In the following analysis, we will consider the probabil-ity distribution of the normal forces during uniaxial com-pression with reference to the compressive z−directionand the two lateral x and y−direction. Keeping each di-rection as reference, we define a cut off χ such that thecontact forces admitted for the probability distributionanalysis fulfill the criteria |nc · nǫ| > χ where nc is thenormal unit vector of the reference direction and nǫ isthe strain eigenvector corresponding to a compressive ortensile direction. The strain eigenvector is fixed due tothe deformation mode, but will be different for other testset-ups. In the case χ = 0, all contact forces in the en-

Page 15: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

15

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Σwsl

/Σsl

τ

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(a)

0.42

0.44

0.46

0.48

0.5

0.52

0.54

0.56

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

Σsst

/Σst

τ

µ=0.01

µ=0.02

µ=0.05

µ=0.1

µ=0.2

µ=0.3

µ=0.5

µ=1.0

(b)

FIG. 12: Fraction of (a) weak sliding contacts (wsl) and (b) strong sticking contacts (sst) with respect to the totalnumber of sliding (

sl) and sticking (∑

st) contacts, respectively, during uniaxial loading- and unloading fordifferent friction coefficients.

semble will be considered while no contacts exist whenχ = 1. For the present study, we set χ = 0.8 and onlynote that as χ approaches 1, less data are available andthe noise level increases, but not changing much the fol-lowing results.

In Fig. 13, we plot the normalized probability densityof the normal force P (f/〈fall〉) against the normalizedforce f∗ = f/〈fall〉 for the three reference directions (x,y and z) and for all contacts. In this case, µ = 0.1 andcut-off χ has been set to 0.8. To allow for comparison,the forces have been normalized with the mean of thenormal force for all contacts. As shown in Fig. 13(a), atτ = 0, the force probabilities from the three referencedirections and for all contacts are virtually the same evi-denced by the apparent collapse of the different curves oneach other. This is not surprising since the initial stateis isotropic and no direction-dependent deformation hastaken place. At maximum compression (in Fig. 13(b)), adifference between the force distribution in the compres-sion z−direction and the axial x and y direction is evi-dent. Firstly, we observe that the force probability in theradial directions (x,y) are close since no active deforma-tion takes place in these directions. Another observationis that the proportion of weak forces in the axial direc-tion significantly exceeds those in the compression direc-tion. Also, the mean force in the z−direction is observedto be higher than in the radial direction. At maximumcompression, we observe a wider distribution in the com-pressive z− direction compared to the force distributionfor all contacts and the two lateral directions.

The longer tail seen in the force distribution in thez−direction is due to the presence of stronger forces com-pared to the other directions. At the end of the decom-pression cycle, shown in Fig. 13(c), we observe that theinitial state is not recovered due to the deformation his-tory of the sample and there is a higher proportion ofweak forces in the decompression z−direction comparedto the radial directions.

V. POLAR REPRESENTATION

To understand the orientation and arrangement of thecontacts over the whole angular spectrum during uniaxialdeformations, we introduce now the polar representationof contacts, forces and mobilized friction. For the analy-sis, we test two different averaging methods, namely theconstant bin width (b) and constant height (h), whichgive comparable results and are shown in detail in Ap-pendix I. In the following, we will use data obtained usingthe constant bin width method.

A. Harmonic approximations

The axial distribution of contact force orientationsP (θ), along with the degree of anisotropy in a granularpacking, can be approximated by a Legendre polynomialbased on spherical harmonics of the form Y m

l (θ,ϕ) [7, 10].The approximation is simplified by admitting only func-tions that are consistent with the symmetry of the defor-mation mode, namely functions independent with respectto ϕ and periodic as a function of θ. With this crite-ria, the two lowest admissible functions are Y 0

0 =1 andY 02 = 3 cos2 θ − 1 such that the second order harmonic

representation of contacts is of the form:

P2(θ) = a0[1 + ǫ(3 cos2 θ − 1)] (23)

with the factor a0 as constant and a unique anisotropydescriptor ǫ. In our case, due to normalization, a0 ≈ 0.5..For the uniaxial mode, snapshots of the contact probabil-ity density data are presented in Fig. 15(a) during uni-axial loading and unloading. We observe distributionswith two peaks and a dip around π/2 indicating that ahigher order approximation is needed. The higher orderneeded for the present uniaxial dataset is possibly due tothe peculiarity of the deformation mode. Unlike the tri-axial test which involves an active stress control on thelateral boundaries of the system, the stress on the lat-eral boundaries of the uniaxial mode evolve, albeit with

Page 16: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

16

0

0.02

0.04

0.06

0.08

0.1

0.12

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

P(f

/<f a

ll>

)

f/<fall>

xyz

All

(a)

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

P(f

/<f a

ll>

)

f/<fall>

xyz

All

(b)

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

P(f

/<f a

ll>

)

f/<fall>

xyz

All

(c)

FIG. 13: Normalized probability density of the normal force P (f/〈fall〉) for the three reference directions and for allcontacts forces plotted against the normalized force f∗ = f/〈fall〉 for µ = 0.1 and cut-off χ = 0.8. Three snapshots

are shown at (a) initial, τ= 0, (b) maximum, τ =0.5, and (c) final, τ= 1.0 compression.

θ P2(θ) P6(θ)

θ = 0 a0[1 + 2ǫ] a0[1 + 2ǫ2 + 8ǫ4 + 16ǫ6]

θ = π/2 a0[1− ǫ] a0[1− ǫ2 + 3ǫ4 − 5ǫ6]

θ = π a0[1 + 2ǫ] a0[1 + 2ǫ2 + 8ǫ4 + 16ǫ6]

TABLE III: Second and sixth order harmonic expansionof the contact distribution for the axial direction(compression: θ = 0, π) and the lateral direction

(θ = π/2).

smaller magnitude in comparison to the stress in the axialdirection.

Eq. (23) can be extended to admit higher order spher-ical harmonic functions with l = 4, 6. For l= 4, Y 0

4 =35 cos4 θ − 30 cos2 θ + 3 and for l =6, Y 0

6 = 231 cos6 θ −315 cos4 θ+105 cos2 θ−5 all with different prefactors. Fora 6th order expansion, the contact distribution will takethe form:

P6(θ) = a0[1 + ǫ2Y02 + ǫ4Y

04 + ǫ6Y

06 ], (24)

where the axial symmetry is implied. Eq. (24) introducesnow three anisotropy state descriptors ǫi, with i = 2, 4, 6.Also, Equations (23) and (24) can be further simplifiedfor the well defined limits at θ = 0, π/2 and π as shownin Table III.

Different methods of obtaining the anisotropy state de-scriptors have been attempted in this study. The de-tails and comparison of the methods are discussed inAppendix II. For all methods, we consistently observethat the contact distribution is approximated by a sixthorder polynomial with two peaks and a strong depressionat π/2. In the following, as a reference case, we use theazimuthal fit to the constant probability data discussedin Appendix II.

B. Discussion of Results

Having established that the contact distribution isapproximated by a sixth order distribution with threeanisotropy state descriptors, we compare descriptors ǫ2,ǫ4 and ǫ6 for different friction coefficients as functions ofthe deviatoric strain during uniaxial loading and unload-ing. From Fig. 14(a), besides a slight increase in the max-imum ǫ2 values between µ = 0 and 0.02, the maximum ǫ2value shows a decreasing trend with friction and almostsaturates for the highest friction coefficients. This is con-sistent with the trend of the maximum deviatoric fabricshown in Fig. 3. Also consistent with the deviatoric fabricevolution during unloading is that the initial state is notrecovered. In Fig. 14(b), beginning from different ran-dom values, ǫ4 is negative and systematically decreasesfor all friction coefficients during loading followed by aslight increase during unloading. The descriptor ǫ6 isdistributed around zero and remains fairly constant dur-ing loading and unloading but has some variation withineither deformation.

In addition to the contact probabilities in Fig. 15(a),we now study the distribution for other quantities. Thepolar distributions of the normal force shown in Fig.15(b) during loading displays a high and increasing nor-mal force along the compression (axial 0◦) direction com-pared to the lateral (π/2) direction reaching their max-imum at τ = 0.5. After strain reversal (unloading), thenormal force in the tensile (extension) direction is re-duced until the force in the lateral (π/2) direction be-comes higher. Interestingly, in contrast to P (θ), the dis-tribution of the normal forces fn(θ) is well described by asecond-order harmonic approximation similar to Eq. (23)during loading and unloading.

For the distribution of the tangential force and mobi-lized friction, shown respectively in Fig. 15(c) and 15(d),we observe a distribution similar to that of the contactsshown in Fig. (17), with two strong peaks and a depres-sion around π/2. This indicates the need for a higherorder tensorial descriptor also for these two quantitiesthat appear to be strongly related with the behavior of

Page 17: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

17

-0.03

-0.02

-0.01

0

0.01

0.02

0.03

0.04

0 0.04 0.08 0.12

∈2

εdev

0.12 0.08 0.04 0

εdev

µ=0.0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.5µ=1.0

(a)

-0.015

-0.01

-0.005

0

0.005

0 0.04 0.08 0.12

∈4

εdev

0.12 0.08 0.04 0

εdev

µ=0.0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.5µ=1.0

(b)

-0.015

-0.01

-0.005

0

0.005

0 0.04 0.08 0.12

∈6

εdev

0.12 0.08 0.04 0

εdev

µ=0.0µ=0.01µ=0.02µ=0.05

µ=0.1µ=0.2µ=0.5µ=1.0

(c)

FIG. 14: Comparison of the 6th order anisotropy statedescriptors (a) ǫ2 (b) ǫ4 (c) ǫ6 as a function of the

deviatoric strain for different friction coefficients duringloading (left) and unloading (right).

the contact network, rather than with the normal forces.Similar to fn(θ), the distribution of the tangential forceft(θ) also shows an increase along the compression direc-tion followed by a decrease during decompression. Wealso find that during loading, the mobilized friction in-creases along the tensile (π/2) direction while remainingfairly stable and flat in the lateral direction. After strainreversal, the mobilized friction increases again along thetensile direction (which is now 0◦). Coupling these obser-vations to the normal force distribution, we find that fric-tion is less mobilized along the direction where strongerforces exist (compression) and more mobilized along thedirection where weaker forces (tension) are seen. Simi-lar to the directional probability distribution of the nor-mal force presented in Section IVD, the initial state (atτ = 0) which is mostly isotropic is not recovered at theend of unloading (τ = 1).

VI. SUMMARY AND OUTLOOK

The discrete element method has been used to inves-tigate the microscopic and macroscopic response of fric-tional, polydisperse granular assemblies under uniaxialloading and unloading paths. The main goal was to inves-tigate the effects of contact friction on the force and con-

tact network orientation and distribution and to relatethis to the evolution of structural anisotropy – which isthe key ingredient that quantifies the response of granularmaterials under non-isotropic loading conditions. Sincethe uniaxial test is widely realizable in laboratory exper-iments using various geometries, our findings should berelevant for both experimental and numerical researchersinterested in the behavior of packings under different de-formation and stress conditions. The present study cov-ers a wide range of friction coefficients for systems thatare already “jammed” and since the boundary walls areperiodic, the effects of walls and system geometry shouldbe minimal, which allows to understand the bulk behav-ior with rather few particles (N ≈ 10,000).

As preparation procedure, in order to obtain homoge-neous initial isotropic states, we attempted several prepa-ration protocols and found that the methods lead tomostly identical initial states. We report that the evo-lution of the deviatoric stress ratio and the deviatoricfabric, as functions of the deviatoric strain, in the pres-ence of friction, are different with respect to each other.Even though the contact model is linear, both quanti-ties show a non-linear behavior due to the structuralchanges during loading and unloading. For the config-urations with lower friction, a saturation in the devia-toric stress ratio during loading was observed. However,when friction is increased, a clear saturation of the devi-atoric stress ratio is not seen within the same maximumstrain. During loading, the deviatoric fabric, which quan-tifies the structural anisotropy reaches a maximum be-fore maximum εdev independent of µ, and then decreasesas the maximum strain is approached. Interestingly forthe higher friction coefficients, a second increase was ob-served. The peak deviatoric stress ratio smax

dev reachedduring uniaxial loading increases up to µ = 0.3 and sub-sequently decreases for higher friction. The peak devi-atoric fabric reached Fmax

dev largely shows a decreasingtrend with increasing friction and eventually saturatesat Fmax

dev ≈ 0.025.

The orientation of the largest stress eigenvector θs1,during loading with non-systematic rates aligns to 0◦,i.e. the compression direction. When strain is reversed,we observe that θs1 remains oriented along the verticaldirection before reverting to 90◦. The deviatoric strainat which the reversal happens is observed to decreasewith increasing friction. On the other hand, the orienta-

tion of the largest fabric eigenvector θf1 shows a strongdependence on friction. For systems with low friction,

θf1 aligns with the compressive direction during loadingwhile the configurations with high friction begin to alignperpendicular to the axial direction as maximum devia-toric strain is approached.

The deviatoric strains at which the stress tensorchanges, i.e. sdev = 0, θs1 = 45◦ and stress shape fac-tor Λσ = 0 (zero), are identical to each other and showa decreasing dependence on friction, with εdev ≈ 0.01 forµ = 1. For quantities relating to the microstructure, e.g.the strain at which e.g. Fdev = 0 is different from the

Page 18: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

18

0.4 0.42 0.44 0.46 0.48 0.5

0.52

P6(

θ)

τ=0.248

0.4 0.42 0.44 0.46 0.48 0.5

0.52 P

6(θ)

τ=0.5

0.4 0.42 0.44 0.46 0.48 0.5

0.52

P6(

θ)

τ=0.748

0.4 0.42 0.44 0.46 0.48 0.5

0.52

P6(

θ)

τ=0.872

0.4 0.42 0.44 0.46 0.48 0.5

0.52

0 π/4 π/2 3π/4 π

P6(

θ)

τ=0τ=1

(a)

6000

10000

14000

18000

f n(θ

)

τ=0.248

6000

10000

14000

18000

f n(θ

)

τ=0.5

6000

10000

14000

18000

f n(θ

)

τ=0.748

6000

10000

14000

18000

f n(θ

)

τ=0.872

6000

10000

14000

18000

0 π/4 π/2 3π/4 π

f n(θ

)

τ=0τ=1

(b)

400

600

800

1000

1200

f t(θ

)

τ=0.248

400

600

800

1000

1200

f t(θ

)

τ=0.5

400

600

800

1000

1200

f t(θ

)

τ=0.748

400

600

800

1000

1200

f t(θ

)

τ=0.872

400

600

800

1000

1200

0 π/4 π/2 3π/4 π

f t(θ

)

τ=0τ=1

(c)

0.55

0.65

0.75

0.85

ψt(θ)

τ=0.248

0.55

0.65

0.75

0.85

ψt(θ)

τ=0.5

0.55

0.65

0.75

0.85

ψt(θ)

τ=0.748

0.55

0.65

0.75

0.85

ψt(θ)

τ=0.872

0.55

0.65

0.75

0.85

0 π/4 π/2 3π/4 π

ψt(θ)

τ=0τ=1

(d)

FIG. 15: Snapshots of the polar distribution of the (a) contacts P6(θ) (b) normal force fn(θ) (c) tangential forceft(θ) (d) mobilized friction ψt(θ) at different dimensionless time (τ) during uniaxial loading and unloading for

friction µ = 0.1.

stress.As reported in previous studies [38], we also confirm

that the orientation of the eigenvector of stress corre-sponding to the compression direction for strong (forcesgreater than average) and weak (forces less than average)contacts are orthogonal with respect to each other. Asa consequence of the definition of the stress tensor, theeffects of strong contacts are more dominant for stress.For fabric however, strong contacts are more dominantwhen friction is low while the weak contacts are found toplay a bigger role for stronger friction.In terms of the proportion of contacts, we confirm that

a larger proportion of the total contacts are weak whilethe proportion of sliding contacts out of the total con-tacts are less than 45%. More importantly, we find lessthan 50% weak sliding contacts with respect to the totalnumber of weak contacts. On the other hand, the pro-portion of weak sliding contacts with respect to the totalsliding contacts are significantly higher. The latter is inagreement with earlier studies that show that friction ismore highly mobilized in weak contacts [44, 45].As a consequence of the isotropic initial configuration,

we find that the directional distribution of normal forcesat the initial state are isotropic for all µ. At maximumcompression, we observe a higher mean, a lower peak anda wider tail of the force distribution in the compressivez−direction while the distribution in the two lateral di-rections remain identical, narrower and have a shortertail. Due to history, after uniaxial compression and ten-sion, the initial states are not recovered with weaker

forces and less contacts in the tensile z− direction.

We have also presented two averaging methods for thepolar representation of contacts using the constant az-imuthal and constant height methods. For our data, asecond order tensor is insufficient to describe the struc-tural anisotropy. We find a sixth order distribution withtwo peaks leading to three anisotropy state descriptors(ǫ2, ǫ4 and ǫ6).

We find that ǫ2 is close to F sdev/

√3 and for different

friction, the maximum ǫ2 values behave in a similar fash-ion to the maximum deviatoric fabric. Finally, whilea second order tensorial descriptor is sufficient for thenormal force, the tangential force and mobilized frictionshow a similar behavior to the contact distribution – re-quiring a higher order harmonic approximation due tothe two strong peaks at π/4 and the dip around π/2.

Future studies will concern exploring higher ordertensors and the validity of the findings for other non-isotropic deformation modes (e.g. under simple and pureshear). Furthermore, recent experiments [9] will allow tovalidate the present observations by DEM.

Appendix I: Averaging Methods

In this appendix, we describe the two averaging meth-ods namely the constant azimuthal angle method (b) andthe constant height method (h).

Page 19: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

19

A. Constant azimuthal angle (bin width) method

Given the three normal unit vector components nx, ny,and nz for each contact pair, to calculate the azimuthalangle, one needs the polar orientation arccos(nz) of thenormal unit vector in the direction relative to the active(axial) direction as schematically described in Fig.16(a).We average over the spherical azimuthal (vs. polar) (r, ϕ)coordinate and then distribute the vectors, based on theirorientation into bins of width ∆θ.The fraction of contacts in a single bin is defined as

φθ = Cθ/Ctot, where Cθ =

C∈b 1 and b ∈ [θ−∆θ/2; θ+

∆θ/2]. Furthermore, φθ is normalized with the surface ofthe spherical annulus for each b by the factor ∆θ sin θ toyield the azimuthal contact probability density P (θ) =(φθ/∆θ sin θ) such that

∫ π

0P (θ) sin θ∆θ = 1. [58] The

polar distributions of the normal forces, tangential forcesand mobilized friction are given respectively, by fn(θ) =(∑

C∈b fn)/(Cθ), ft(θ) = (

C∈b ft)/(Cθ) and ψt(θ) =

(∑

C∈b(ft/µfn))/(Cθ), where the normalization with the

number of contacts in each bin has been used.

(a) (b)

FIG. 16: Schematic representation of the angles of thecontact unit vector (green arrow) for (a) the constantbin width ∆θ method and (b) the constant height

method. The angles θ and ϕ are the azimuthal angleand polar angle of the system, respectively.

B. Constant height method

In the constant height method, we sort the vectorsbased on their orientations into n azimuthal sphericalsegments with equal heights ∆h = cos θ2−cos θ1 as shownschematically in Fig 16(b). Given the polar radius r, andthe height from the center of each segment to the middleof the sphere h, the polar angle θ of each vector is calcu-lated for every nz ∈ h. The fraction of contacts withineach segment range is then given as φh = Ch/Ctot, whereCh =

C∈h 1 and h ∈ [h − ∆h/2;h + ∆h/2]. With−1 ≤ h ≤ 1, specifying the number of bins Mh (e.g.Mh = 20), allows to compute all h−intervals and bound-aries.Other quantities, including the normal and tangential

forces and mobilized friction can be computed similar to

the constant bin width method, just by summation andnormalization with Ch instead of Cθ.

Appendix II: Fit Methods

In the following, we describe different methods of ob-taining the anisotropy state descriptors using the dataobtained using the constant bin width or the constantheight methods.

C. Method 1: Fit azimuthal contact probabilitydensity P (θ) = (φθ/∆θ sin θ)

In the first case, we fit the azimuthal contact proba-bility density data P (θ) = (φθ/∆θ sin θ) using the har-monic equation (24). Note that for the special case ofuniaxial compression, Eq. (23) does not lead to consis-tent results across the methods and is thus disregarded.However, as shown in different literature especially un-der triaxial compression, the second order approximationP2(θ) is sufficient to fully capture the contact probabilitydensity data. Exemplary (not exhaustive) references ofworks where forms of Eq. (23) as applied to various tri-axial tests are presented in table IV. Other experimentaland numerical set-ups are also shown. Note that in thesecases, the orientational contact distribution obtained isnot of second order and are not fitted.

D. Method 2: Fit to the constant height data

In the second case, we directly fit the fraction of con-tacts φh, generated using the constant height methodsuch that the bad statistics at the poles are not over-exposed as in Method 1. In this case, we set the zeroorder parameter a0 = 0.5.

0.45

0.46

0.47

0.48

0.49

0.5

0.51

0.52

0.53

0.54

0 0.5 1 1.5 2 2.5 3 3.5

P6(

θ)

bM1M2M3h

(a)

FIG. 17: 6th Order Fits of methods 1 and 2 and 3 tothe P (θ) = (φθ/∆θ sin θ) data at dimensionless timeτ = 0.076. The solid red symbols represent data

obtained using the constant height method (h), whilethe triangles are those obtained with the constant bin

width method (b).

Page 20: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

20

Reference Mode Contact Harmonic function

probability P (θ)

1. Azema et al. [7] triaxial second order 14π

[1 + ǫ(3 cos2 θ − 1)]

2. Deng and Dave [11] particle Settling higher order –

in a cylindrical geometry

3. Ishibashi et al [18] triaxial second order K(1− ǫ+ 3ǫ cos2 θ)

4. Jenkins [19] triaxial second order C4π

[(1− ǫ) + 3ǫ cos2 θ]

5. Silbert et al. [43] 3D particle settling higher order –

on a flat base

6. Staron and Radjai [44] 2D avalanche higher order –

TABLE IV: Selected references on the orientational contact distribution for various modes. For an isotropic sample,K = C/4π and C is the coordination number.

E. Method 3: Fit to fraction of contacts φθ with a∆θ sin θ scaling

In the third case, we fit the fraction of contacts φθ datadirectly using harmonic equation (23) or (24) multipliedby ∆θ sin θ. The original signal is a first order sinus, i.e.less weighting is given to the areas close to the poles suchthat their larger statistical errors are suppressed.In Fig. 17, we show the sixth order harmonic fits using

methods 1 (M1), 2 (M2) and 3 (M3) to the constant binwidth (b) and constant height (h) data for µ = 0. Foreach method, three anisotropy state descriptors, namelyǫ2, ǫ4 and ǫ6 are obtained. Here, as an example, we showa single snapshot, namely at τ = 0.076. Note that theoriginal data is from 0 to π/2 and the extension from π/2to π is only a mirror image. Focusing on the numericaldata (symbols), we observe two strong peaks at aboutπ/2± π/4 and a local maximum at π/2. The twin peaksindicate that a distribution higher than second order isneeded. The data are well captured by a sixth orderapproximation P6(θ) (solid lines). Comparing the b andh, we observe stronger scatter at the boundaries for the bdata due to the weak statistics at the extreme θ values (0and π). For the fits, we observe that M1, M2 ad M3 areclose and the major differences between them are mostpronounced at both peaks and extrema.In Fig. 18, we plot the evolution of the anisotropy

state descriptors ǫ2, ǫ4 and ǫ6 as functions of the devia-toric strain and compare the three methods. From Fig.18(a), during loading, ǫ2 grows and reaches a maximumat εdev ≈ 0.025 from where it slightly decreases. Aftermaximum loading, ǫ2 decreases (taking well into account

the sign change) and becomes increasingly negative un-til it reaches ǫ2 ≈ −0.055 at complete unloading (τ=1).Comparing the three methods, M3 is slightly off (higher)during loading while M1 is also slightly off at the end ofunloading. Interestingly, we find that the evolution of ǫ2is similar to the simple definition of the deviatoric fabricin Eq. (19) involving a difference between the fabric com-ponent of the axial direction and the average of the com-ponents of the two lateral components. Note that the def-inition of the fabric used here is based on Eq. (16) whichconsiders only the contacts and not the dependence onthe volume fraction. We note that the magnitude of ǫ2 isproportional to Fdev/

√3 (black diamonds in Fig. 18(a))

with M3 slightly off during loading and the unloadingdata also showing slight variations. From Fig. 18(b) and18(c), the values of ǫ4 and ǫ6 appear small compared toǫ2 but must not be neglected. When ǫ2 is taken as thestructural anisotropy state descriptor, it much resemblesFdev. However, the higher order anisotropy is quantifiedby the ǫ4 (which is negative) and ǫ6 (which is stronglyfluctuating and different for different methods).

Acknowledgement

Helpful discussions with N. Kumar, A. Singh, V. Og-arko, F. Radjai and J.N. Roux are gratefully acknowl-edged. This work is financially supported by the Euro-pean Union funded Marie Curie Initial Training Network,FP7 (ITN-238577), see http://www.pardem.eu/ for moreinformation.

[1] F. Alonso-Marroquin, H. J. Herrmann. Phys. Rev. Let.,92:054301, 2004.

[2] X. Z. An, R. Y. Yang, R. P. Zou, and A. B. Yu Powd.Tech., 188(2):102–109, 2008.

[3] S. J. Antony and N. P. Kruyt. Phys. Rev. E.,

79:031308(3), 2009.[4] S. J. Antony and M. A. Sultan. Phys. Rev. E.,

75:031307(3), 2007.[5] E. Azema and F. Radjai. Phys. Rev. E., 81:051304:1–17,

2010.

Page 21: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

21

-0.06

-0.05

-0.04

-0.03

-0.02

-0.01

0

0.01

0.02

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

∈2

εdev

M1M2M3Fdev/√ 3

(a)

-0.01

-0.008

-0.006

-0.004

-0.002

0

0.002

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

∈4

εdev

M1M2M3

(b)

-0.01

-0.008

-0.006

-0.004

-0.002

0

0.002

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16

∈6

εdev

M1M2M3

(c)

FIG. 18: Evolution of the anisotropy state descriptors ǫ2, ǫ4 and ǫ6 of the 6th order expansion Eq. (24) as functionof deviatoric strain using the three methods for µ = 0.

[6] E. Azema, F. Radjai, R. Peyroux, and G. Saussine. Phys.Rev. E., 76:011301, 2007.

[7] E. Azema, F. Radjai, and G. Saussine. Mech. Mat.,41(6):729–741, 2009.

[8] M. M. Bandi, M. K. Rivera, F. Krzakala and R. E. Ecke.Phys. Rev. E., 87:042205, 2013.

[9] N. Brodu, J. Dijksman and R. P. Behringer. PrivateCommunication, 2013.

[10] H. J. Bunge and P. R. Morris. Butterworths London,1982.

[11] X. L. Deng and R. N. Dave. Gran. Mat., 15:401–415,2013.

[12] P. J. Digby. J. Appl. Mech., 48(4):803–808, 1981.[13] A. Ezaoui and H. Di Benedetto. Geotechnique, 59(7):621–

635, 2009.[14] F. Goncu, O. Duran, and S. Luding. C. R. Mecanique,

338(10-11):570–586, 2010.[15] N. Guo and J. Zhao. Comp. Geotech., 47:1–15, 2013.[16] O. I. Imole, N. Kumar, V. Magnanimo, and S. Luding.

KONA Powd. Part. J., 30:84–108, 2013.[17] O. I. Imole, M. Wojtkowski, V. Magnanimo, and S. Lud-

ing. Powd. & Grains vol. 1542, 601–604, 2013.[18] I. Ishibashi, T. Agarwal, and S. A. Ashraf. Intl. Conf.

1st DEM vol. 23, 1989.[19] J. T. Jenkins, Micromech. Gran. Mat., 245-251, 1987.[20] N. Kumar, O. I. Imole, V. Magnanimo, and S. Luding.

Particuology, Accepted, 2013.[21] N. Kumar, O. I. Imole, V. Magnanimo, and S. Luding.

Sydney, Australia, 2013. Balkema.[22] N. Kumar, V. Magnanimo, and S. Luding. Acta Geotech-

nica, Under Review, 2013.[23] A. Kwade, D. Schulze, and J. Schwedes. Powd. hand. &

Proc., 6(1):61–65, 1994.[24] A. Kwade, D. Schulze, and J. Schwedes. Powd. hand. &

Proc., 6(2):199–203, 1994.[25] L. La Ragione and V. Magnanimo. Phys. Rev. E.,

85,031304, 2012.[26] S. Luding. Int. J. Sol. & Struct., 41(21):5821–5836, 2004.[27] S. Luding. J. Phys. Cond. Mat., 17(24):S2623–S2640,

2005.[28] S. Luding. Gran. Mat., 10(4):235–246, 2008.[29] S. Luding. Part. Sci. and Tech., 26(1), 2008.[30] S. Luding and E. Bauer. Geomech. & Geotech., 495–499.

CRC Press/Balkema, 2011.[31] M. Madadi, O. Tsoungui, M. Latzel, and S. Luding. Int.

J. Sol. & Struct., 41(9-10):2580, 2004.[32] V. Magnanimo, L. La Ragione, J. T. Jenkins, P. Wang,

and H. A. Makse. Euro. Phys. Let., 34006, 2008.[33] T. S. Majmudar and R. P. Behringer. Nature,

435(7045):1079–1082, 2005.[34] GDR MiDi. Eur. Phys. J. E., 14(4):341–365, 2004.[35] M. Oda. Soils & Found., 1(12):17–36, 1972.[36] V. Ogarko and S. Luding. J. Chem. Phy., 136(12), 2012.[37] F. Radjaı, S. Roux, and J. J. Moreau. Chaos, 9(3):544–

550, 1999.[38] F. Radjai, D. E. Wolf, M. Jean, and J. J. Moreau. Phys.

Rev. Let., 80:61–64, 1998.[39] M. Satake. Proc., IUTAM Symp. on Deformation and

Failure of Granular materials, Delft, The Netherlands.1982.

[40] J. Schwedes. Gran. Mat., 5(1):1–43, 2003.[41] M. R. Shaebani, M. Madadi, S. Luding, and D. E. Wolf.

Phys. Rev. E, 85, 011301, 2012.[42] L. E. Silbert, D. Ertas, G. S. Grest, T. C. Halsey, and

D. Levine. Phys. Rev. E, 65, 031304, 2002.[43] L. E. Silbert, G. S. Grest, and J. W. Landry. Phy. Rev.

E., 66:061303(6 Part 1), 2002.[44] L. Staron and F. Radjai. Phys. Rev. E., 72:041308:1–5,

2005.[45] L. Staron, F. Radjai, and J. Vilotte. J. Stat. Mech.,

2006(07):P07014:1–15, 2006.[46] A. S. J. Suiker and N. A. Fleck. J. Appl. Mech., 71:350–

358, 2004.[47] C. Thornton and S. J. Anthony. Phil. Trans. Roy.

Soc. Lond.: Ser. A: Math., Phys. and Eng. Sci.,356(1747):2763–2782, 1998.

[48] C. Thornton and L. Zhang. Phil. Mag., 86(21-22):3425–3452, 2006.

[49] R. Tykhoniuk, J. Tomas, S. Luding, M. Kappl, L. Heim,and H. J. Butt. Chem. Eng. Sci., 62(11):2843–2864, 2007.

[50] M. van Hecke. J. Phys. Cond. Mat., 22:033101(3), 2009.[51] K. Walton. J. Mech. and Phys. Sol., 35(2):213–226, 1987.[52] T. Weinhart, R. Hartkamp, A. R. Thornton, and S. Lud-

ing. Phys. of Fluids, 25:070605(7), 2013.[53] T. Weinhart, A. R. Thornton, S. Luding, and

O. Bokhove. Gran. Mat., 14(2):289–294, 2012.[54] C. M. Wensrich, E. H. Kisi, J. F. Zhang, and O. Kirstein.

Gran. Mat., 14(6):671–680, 2012.[55] H. Zetzener and J. Schwedes. 6th Intl. Conf. on Handling

and Transportation, pages 51–55, 1998.[56] J. Zhang, T. Majmudar, A. Tordesillas, and

R. Behringer. Gran. Mat., 12(2):159–172, 2010.[57] An alternative way to enforce the sign convention is to

multiply the deviatoric stress Eq. (14) by the sign of the

Page 22: NOT FOR DISTRIBUTION REVIEW COPY - TFE1 | Homepharmaceutical, agriculture and mining – among others. In many granular systems, interesting phenomena like dilatancy, anisotropy, shear-band

22

difference between the eigenvalue of the main compres-sive direction and the average in the other two fixed direc-tions as given for fabric in Eq. (19). This leads to positiveand negative sdev, which should take care of the strainreversal [21, 22].

[58] An alternative to the ∆θ sin θ normalization is a discreteformulation cos θf − cos θi where θi = θ−∆θ/2 and θf =θ +∆θ/2.


Recommended