You are on page 1of 15

HOSTED BY Available online at www.sciencedirect.

com

ScienceDirect
Soils and Foundations xxx (2018) xxx–xxx
www.elsevier.com/locate/sandf

Induced fabric anisotropy of granular materials in biaxial tests


along imposed strain paths
Jingshan Shi, Peijun Guo ⇑
Department of Civil Engineering, McMaster University, Hamilton, Ontario L8S 4L7, Canada

Received 15 February 2017; received in revised form 7 November 2017; accepted 23 November 2017

Abstract

In granular materials, loading on an initially isotopic assembly usually induces particle rearrangement, and this is referred to as the
induced anisotropy of fabric. A series of biaxial tests are conducted along various strain paths using DEM to investigate the evolution of
the induced anisotropy. The evolution of both the overall contact network and the sub-networks (the strong and weak) are examined
separately. Results of DEM simulations indicate that the evolution of the fabric deviator in the overall contact network can be described
as a power function of the stress ratio prior to the peak stress ratio that depends on the imposed dilation rate. A unique fabric-stress
relation is obtained for the strong sub-network, which is independent of the strain path, the initial porosity and the confining pressure.
Moreover, deformation instability is observed only along dilatant strain paths, which can be related to the degradation or even collapse
of a weak sub-network, even though the strong sub-network dominates the strength of the granular assembly.
Ó 2018 Production and hosting by Elsevier B.V. on behalf of The Japanese Geotechnical Society.

Keywords: Granular material; Fabric evolution; Contact network; Imposed strain paths

1. Introduction soils, including dilation (i.e. shear-induced volume change),


failure and instability during deformation (Fu and
The mechanical behavior of a granular material can be Dafalias, 2011; Muhunthan and Collins, 2003; Oda, 1993;
inherently anisotropic as a consequence of the original Radjaı̈ et al., 2012; Wan et al., 2007; Wan and Pouragha,
manner in which the material was formed or deposited. 2015; Wan and Guo, 2004). Research incorporating aniso-
Such anisotropy is easily altered when the granular mate- tropy into the constitutive model has been conducted by
rial is subjected to shear distortion, because of the rear- Guo and Stolle (2005), Lade (2008), Li and Dafalias
rangement of particles via relative particle movements, (2011) and Gao et al. (2014) among others. Yet, one of
including rolling and sliding at particle contacts. The spa- the obstacles of micromechanically formulated constitutive
tial arrangement and inter-connectivity of discrete particles laws lies in how to mathematically describe the anisotropy
subjected to stress change is also known as stress-induced and its evolution with the deformation history. For granu-
anisotropy (Oda, 1993). Both experimental and theoretical lar materials, geometrical anisotropy can be quantified by a
studies suggest that induced anisotropy and its evolution fabric tensor based on various direction-dependent mea-
contribute to key aspects of the mechanics of granular sures (Kanatani, 1984). The fabric tensor defined by
Satake (1978) in terms of the contact normal orientation
is expressed as /ij ¼ hni nj i, where ni is the i-th component
Peer review under responsibility of The Japanese Geotechnical Society.
⇑ Corresponding author. of contact normal vector and hAi denotes the average of
E-mail addresses: shij9@mcmaster.ca (J. Shi), guop@mcmaster.ca quantity A over all N c contacts within the granular
(P. Guo). assembly.

https://doi.org/10.1016/j.sandf.2018.02.001
0038-0806/Ó 2018 Production and hosting by Elsevier B.V. on behalf of The Japanese Geotechnical Society.

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
2 J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx

Various models have been developed to quantify the rise via buckling to an indirect anisotropy inside the dissi-
fabric evolution or induced anisotropy characterized by pative weak network with a preferred direction orthogonal
the evolution law of fabric in terms of applied stresses to the major principal direction of the stress tensor. Intu-
(Fu and Dafalias, 2015; Mehrabadi and Nemat-Nasser, itively, one expects that the strong and weak sub-
1983; Nakai, 1989; Oda, 1993). Results of biaxial tests on networks play different roles in the deformation and failure
a stack of photo-elastic rods (Satake, 1987) demonstrate process of granular materials.
that the ratio of principal stresses (r1 =r2 ) is approximately Tests along proportional strain paths have been consid-
proportional to the square root of the corresponding ratio ered more suitable for the investigation of deformation
of the principal values of the fabric tensor (/1 =/2 ) such as instability of granular materials since the difference
1=2 between the imposed rate of volume change and the inher-
ðr1 =r2 Þ  ð/1 =/2 Þ . The same conclusion is also obtained
by Maeda et al. (2006). By defining the stress tensor and ent potential of dilation determines whether or not the
fabric tensor as rmn ¼ N c^l0 hf m nn i and F mn ¼ N c^l0 hnm nn i, material succumbs to unstable deformation (Guo and Su,
Oda et al. (1982) obtain a relation between their principal 2007). In addition, during these tests, instability occurs
components ri and F i (i ¼ 1; 2) in the form of strictly inside the conventional plastic limit condition and
corresponds to the vanishing of the second-order work
ri ¼ a0 F i þ b0 F 2i , where a0 ; b0 are material constants with
(Prunier et al., 2009). These tests can be found, for exam-
f and ^l0 being the contact force vector and the average ple, in Darve and Laouafa (2000), Guo and Su (2007),
length of the branch vector, respectively. This relation is Nicot et al. (2015, 2013), Prunier et al. (2009) and Wan
verified using the results from biaxial compression tests et al. (2005). By conducting biaxial tests along proportional
on assembly of photo-elastic rods (Oda et al., 1982). In strain paths on assemblies of photo-elastic particles, Wan
DEM simulations by Ng (2001) for 3D compression tests et al. (2005) observe that the lateral confinement (the weak
of ellipsoidal arrays, the strength indicator defined as force columns) indicates the strength of central core (strong
b1 ¼ lnð/1 =/3 Þ is found to be directly proportional to the force columns) which can locally collapse, and hence leads
principal stress ratio ðr1 =r3 Þ. Under triaxial stress condi- to failure at the macroscopic level. They also note that
tions, Wan and Guo (2004) assume that the rate of change while the strong force columns in the central core of spec-
of the fabric tensor components is proportional to change imen are critical for the material stability, the evolution of
in the deviatoric stress ratio. Therefore, at a critical state, these strong and weak force columns is hard to trace in lab-
all components of the fabric tensor eventually achieve con- oratory tests. One expects that some in-depth observation
stant values. By conducting 3D drained and undrained towards the relation between deformation instability and
tests using DEM simulations, Zhao and Guo (2013) obtain fabric evolution may be obtained by conducting DEM sim-
a unique relation at critical states between the critical fabric ulations along proportional strain paths.
anisotropy (acr c ) and the mean effective stress (p) as The objective of this paper is to examine the evolution of
acr
c ¼ a C  k a lnðp=p a Þ, where aC is the fabric anisotropy fabric and its correlation with the deformation instability
when p ¼ pa and ka depends on the loading paths. of granular materials sheared along proportional strain
When deformation and shear resistance are concerned, paths. The evolution of fabric for different contact net-
not all contacts play the same role. According to Radjaı̈ works (i.e., the overall contact network, the strong and
et al. (1996), the overall contact network can be separated weak sub-networks) is traced in biaxial tests along various
into a sub-network of ‘‘strong contacts” with normal forces strain paths using DEM. The results reveal that the devia-
larger than the average normal force hf n i, and a sub- tor of the fabric tensor based on the strong sub-network
network of ‘‘weak contacts” with normal forces smaller shows a unique relation with the deviatoric stress ratio,
than hf n i. The strong sub-network is the ‘‘loading- particularly for specimens along proportional strain paths
bearing” network in which the contacts are non-sliding, corresponding to forced dilation. The maximum value of
whereas the weak network is the dissipative network in the deviator of fabric tensor varies with the applied strain
which dissipation takes place due to sliding at contacts ratio, or the maximum dilatancy rate in stress-controlled
inside this network (Radjaı̈ et al., 1999). They also observe tests. It is confirmed that the strong sub-network plays
that the strong sub-network carries the whole deviatoric the primary role in shear resistance, while the weak sub-
load while the weak network contributes to the hydrostatic network and its evolution are significantly affected by the
pressure only. Furthermore, based on the results from imposed dilatancy rate. The examination of the second-
biaxial DEM simulations on an assembly of oval particu- order work shows that the instability of deformation is a
lates, Antony et al. (2004) propose a relation, consequence of the internal structure collapse of the weak
1=2
ðq=pÞ  ð1=2Þð/s22 =/s11 Þ , between deviatoric stress ratio sub-network, which is associated with a decrease in the fab-
(q=p) and the fabric ratio (/s22 =/s11 ) for the strong sub- ric anisotropy of the strong sub-network.
network. Kuhn et al. (2015) note that the strong-contact It should be noted that no attempt has been made to
measures of fabric more closely follow an increase in the reproduce any laboratory test results in this paper. It is
stress ratio than the measure that includes all contacts. In the microscopic mechanism and physical aspects of the
addition, the load-bearing strong sub-network carries a fundamental constitutive features of granular material
direct geometrical anisotropy induced by shear, but it gives observed in the laboratory experimental tests that are of

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx 3

special interest in this research. Biaxial DEM simulations Nc c


r¼ hf  lc i ð3Þ
are carried out in order to capture these microstructural V
aspects and can be extended to 3D simulations for a better where V is the volume of the assembly. The contact force f c
understanding of the catastrophic failures that occur in real includes a normal component and a tangential component
cases.
as f c ¼ f c;n nc þ f c;t tc . Here, nc is the outward unit normal
vector of the contact c aligned with the contact force com-
2. Quantification of microstructure ponent of f c;n , tc is the unit tangent vector aligned with the
tangential component of contact force f c;t . The branch vec-
The measures for fabric anisotropy developed to date tor is expressed as lc ¼ lc mc in which lc and mc are the
can be classified into the following four categories: the pre- length and direction of the branch vector, respectively.
ferred orientations of the particle body, the particle sur- In the following sections, the convention in soil mechan-
face, the contact normal and the void space (Kuhn et al., ics is employed to describe the stress and deformation
2015). Among these measures, the most apparent and effi- states. Under biaxial conditions, the following quantities
cient measure associated with the mechanical response of with i, j = 1, 2 are used in later sections: s ¼ p ¼ rii =2,
granular materials is based on the orientation of the con-
t ¼ ðsij sij =2Þ1=2 ¼ ðr1  r2 Þ=2, ev ¼ eii , c ¼ ð2eij eij Þ1=2 ,
tact normal. The fabric tensor quantifying the contact nor-
sin /m ¼ t=s, sin wm ¼ _epv =_cp . In these expressions, s and
mal orientation of the overall contact network is defined as
sij are the spherical and deviatoric parts of the Cauchy
(Satake, 1978)
stress tensor, and ev and eij are the volumetric strain and
1 X Nc
the deviatoric strain tensor. e_ pv and c_ p are the plastic volu-
/ij ¼ nk nk ð1Þ
N c k¼1 i j metric and shear strain increments, respectively. The angles
um and wm are referred to as the mobilized friction angle
where nki is the i-th component of the unit vector represent- and the mobilized dilation angle, respectively.
ing the contact normal at the k-th contact.
Following Radjaı̈ et al. (1996), we evaluate geometrical 3. DEM simulation of biaxial tests along proportional strain
anisotropy of a granular assembly by determining the paths
fabric tensor for the network of all contacts and for sub-
networks that carry normal forces lower (weak sub- According to Wan et al. (2007) and Guo and Su (2007),
network) or higher (strong sub-network) than the average. the strain softening and material instability for granular
By following the definition of the fabric tensor in Eq. (1), materials are dominated by the rate of forced volume
the fabric tensors for the strong and weak sub-networks change and the intrinsic potential of dilation. From a kine-
are expressed as matic point of view, the collapse of a specimen can be illus-
trated as a larger imposed dilatancy rate than the intrinsic
1 X 1 X
Ns
c Nwc

/sij ¼ s nki nkj ; /wij ¼ w nk nk ð2Þ dilatancy rate of the granular assembly. Compared to tests
N c k¼1 N c k¼1 i j along selected stress paths, controlled proportional strain
path tests can provide in-depth observation towards the
where the superscripts s and w stand for quantities associ-
mechanism of strain-softening and deformation instability
ated with the strong and weak contact networks, respec-
of a granular material. As such, DEM simulations for biax-
tively. N sc and N wc are the numbers of contacts in the
ial tests along proportional strain paths are carried out to
strong and weak sub-networks, respectively. The total
investigate the connection between deformation instability
number of contacts Nc in the granular assembly is
and fabric in this study.
N c ¼ N sc þ N wc . By introducing the relative proportion of
It should be noted that the selections of the contact stiff-
the strong contacts in the system, a ¼ N sc =N c , the relation
ness and the inter-particle friction coefficient l have a
between the fabric tensor /ij of the overall contact network
direct influence on the results of DEM simulations. In gen-
and those of the strong and weak sub-networks can be eral, both the Young’s modulus and the Poisson’s ratio of
expressed as /ij ¼ ð1  aÞ/wij þ a/sij . For biaxial cases, the the bulk material are intensively affected by the tangential
degree of fabric anisotropy for different contact networks to normal stiffness ratio v ¼ k s =k n . The micromechanical
within a granular assembly can be characterized by the cor- analyses based on different hypotheses of homogenization
responding fabric deviators, q/ ¼ /1  /2 , qs/ ¼ /s1  /s2 reveal that the Poisson’s ratio of a 3D granular assembly
and qw/ ¼ /w1  /w2 , respectively. It should be noted that composed of equal sized particles can be related to
/ii ¼ 1 holds true in all contact networks according to v ¼ k s =k n via v ¼ ð1  vÞ=ð4 þ vÞ for the kinematic hypoth-
Eqs. (1) and (2). esis (Bathurst and Rothenburg, 1988; Chang et al., 1995)
The stress tensor in a granular medium is a volumetric and v ¼ ð1  vÞ=ð2 þ 3vÞ for the static hypothesis, respec-
averaging of dyadic product of contact force and branch tively. The unrealistic result of v ¼ 0 when v ¼ 1 is likely
vector on all contacts within the representative element vol- owing to the assumption of equal-sized particles. Accord-
ume (REV) (see, e.g., Christoffersen et al., 1981; Cundall ing to a series of DEM simulations, Mohamed and
and Strack, 1979; Rothenburg and Selvadurai, 1981): Gutierrez (2010) find that k s =k n ¼ 1:0 gives a Poisson’s

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
4 J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx

ratio of 0.25, which is realistic and in good agreement with ε11


experimental results. Consequently, the assumption W

k s =k n ¼ 1:0 has been adopted in much of the research 1-1

2-2
involving DEM simulations (Antony et al., 2004; Nguyen
et al., 2009; Zhu et al., 2016). Recent studies reveal that
ε 22

H
the inter-particle friction could influence the critical state
and the stress-dilatancy relationship. According to Huang
et al. (2014), when l < 0.5, the dilatancy tends to increase 1-2

2-1
with the increase of l. However, the behaviour of granular
material at the critical state seems to be insensitive to fur- (a) (b)
ther increases in l when l  0.5. Moreover, high friction
coefficient could result in negative Poisson’s ratio, espe- Fig. 1. (a) Schematics of the specimen, and (b) the associated loading
cially when l ! 1 (Rothenburg and Kruyt, 2004). boundaries in a 2D test.
l ¼ 0:5 has been widely adopted for DEM simulations;
e.g., by Kruyt (2012), Rothenburg and Kruyt (2004) and
Jiang et al. (2006). Referring to Suiker and Chang (2004) and Christoffersen
In this study, a rectangle packing of polydisperse disks is et al. (1981), the uniform strain field eij in a granular assem-
considered as the DEM model, which is constrained by four bly is the best fit for describing the actual contact displace-
frictionless rigid sidewalls. The normal contact stiffness at ments and can be considered as a plausible assumption.
the walls is chosen to be sufficiently high, and the tangential The macroscopic strain rates in the assembly are related
contact stiffness at the walls is set to zero. A linear force- to the relative velocities of the boundary walls as
displacement contact law is employed where the contact vt11 ¼ vt12 ¼ 12 e_ 11  H t and vt21 ¼ vt22 ¼ 12 e_ 22  W t with
behavior is governed by the normal stiffness kn, tangential H t and W t being the height and width of the specimen at
stiffness ks and the friction coefficient l. Here we adopt the time t. Similar methods can be found in Thornton (2000)
assumptions that k s =k n ¼ 1:0 and l ¼ 0:5. The basic charac- as well as Suiker and Chang (2004). The vertical strain rate
teristics of the initial assembly are given in Table 1. e_ 11 is always positive (compression), while the lateral strain
The initial packing of the assembly for each test was rate e_ 22 ¼ _e11 =R can be positive or negative depending on
generated by filling the box with randomly generated par- the select R value. Since R R ð1; 0Þ and the boundary
ticles and then the specimen was slowly consolidated up walls are frictionless, the directions of the major principal
to an initial hydrostatic confining pressure, as shown in stress and strain are always vertical with e_ 11 ¼ e_ 1 and
Fig. 1(a). A total of nineteen numerical tests were con- r11 ¼ r1 . Generally, the major P and minor principal
P strains
ducted. They can be divided into two categories: tests along can be calculated as e1 ¼ e_ 1 Dt and e2 ¼ e_ 2 Dt respec-
proportional strain paths with a constant strain increment tively. Here Dt is the length of time within a step.
ratio R ¼ _e11 =_e22 and tests with constant horizontal stress In all DEM simulations, the velocities of the boundary
that was achieved by controlling the movement of the ver- walls are slow enough (e.g., e_ 1 = 105/s) to maintain a
tical and horizontal walls (as illustrated in Fig. 1b). quasi-static condition throughout the entire loading his-
In this study, the average macroscopic strain is deter- tory. We introduce the parameter # ¼ e_ v =_c, which is the
mined  according to the displacement gradient as ratio of volumetric strain rate (_ev ¼ e_ 1 þ e_ 2 ) to the shear
eij ¼ 12 @ui =@xj þ @uj =@xi . The strain rates in the vertical strain rate (_c ¼ e_ 1  e_ 2 ) and can be related to the strain
(_e11 ) and horizontal (_e22 ) directions are kept constant. ratio R as # ¼ ðR  1Þ=ðR þ 1Þ. Hence, all proportional
strain paths can be categorized into forced dilation with
# < 0 (or 0 < R < 1) and forced contraction # > 0 (or
jRj > 1), respectively. Hereby, a value of # ¼ 0 (or
Table 1
DEM parameters and material properties.
R ¼ 1) is associated with the isochoric deformation condi-
tion, while the hydrostatic compression and K0-
Parameter Value
compression conditions are obtained when # ! 1
Number of particles 26,000 (R ¼ 1) and # ¼ 1 (R ! 1). It should be noted that
Particle density 2000 kg/m2
Smallest particle diameter Dmin 1.6 cm
for initially isotropic specimens, the strain paths with
Largest particle diameter Dmax 2.0 cm 1 < R < 0 are the same as those with R < 1 that
Average particle diameter D 1.8 cm belongs to a subgroup of tests with jRj > 1.
Initial porosity/void ratio 0.14–0.22/0.16–0.28 For the sake of convenience, each individual test along
Inter-particle friction coefficient l 0.5 imposed strain path is described according to the following
Particle-wall friction coefficient l 0.0
Height (H), width (W) of sample H/D = 212, D/W = 110
conventions: (a) the first letter V stands for vertical com-
Initial axial stress r11 170–400 kPa pression; (b) the second letter, C or E, refers to the horizon-
Initial confining stress r22 170–400 kPa tal compression or horizontal extension; and (c) the third
Normal contact stiffness k n k n =D ¼ 2800 MPa letter, C, I or D, stands for volumetric contraction, iso-
Tangential contact stiffness k s k s =D ¼ 2800 MPa choric deformation or volumetric dilation, respectively.

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx 5

Table 2
Details about tests.
Test type Test label Initial stresses Macroscopic boundary conditions
r2 ¼ constant VP200 r1 ¼ r2 ¼ 200 kPa e_ 1 > 0; r2 ¼ 200 kPa
VP300 r1 ¼ r2 ¼ 300 kPa e_ 1 > 0; r2 ¼ 300 kPa
VP400 r1 ¼ r2 ¼ 400 kPa e_ 1 > 0; r2 ¼ 400 kPa
Controlled strain paths VED r1 ¼ r2 ¼ 170 kPa e_ 1 > 0; e_ 2 < 0; j_e1 j < j_e2 j
VEI e_ 1 > 0; e_ 2 < 0; j_e1 j ¼ j_e2 j
VEC e_ 1 > 0; e_ 2 < 0; j_e1 j > j_e2 j
VCC e_ 1 > 0; e_ 2 > 0; j_e1 j > j_e2 j
Note: The sign convention for strain and strain increment is positive for compression.

response changes from stable (hardening) in VEC and


VCC tests corresponding to forced contraction (# > 0) to
300 unstable (strain softening) in VED tests with # < 0. More
PLS
CSL specifically, for the high dilative strain path (i.e.,
VEC
R ¼ 0:43), shearing induces a continuous decrease in the
mean effective stress. The deviator stress first increases
200 VEI and then decreases until ‘‘flow” failure occurs, as if the
specimen were a saturated loose sand subjected to
t, kPa

VCC undrained shearing. For the moderate dilative strain path


100 VED
with R ¼ 0:76, the mean effective stress first decreases while
the deviator stress increases at the beginning of shearing.
Then the effective stress path changes direction with both
s and t increasing with axial strain, which reflects typical
0 strain hardening. Nevertheless, the effective stress path
0 200 400 600 800 1000 snaps back shortly after the maximum friction angle at
s, kPa
(t/s)max is mobilized. The effective stress paths along low
Fig. 2. Stress path responses for tests along various strain paths. dilative strain paths with R ¼ 0:86–2:2 have the same fea-
ture as a saturated dense sand subjected to undrained
shear: the mobilized friction angle (or the t/s ratio) gradu-
For example, VEI represents a test in which the specimen
ally increases to its peak value and then tends to approach
has vertical compression, horizontal extension with the vol-
the critical state, while the mean effective and the deviator
ume being a constant (R ¼ 1:0). For tests with constant
stresses both increase monotonically. For VED tests (# < 0
confining pressure, the second letter P stands for the con-
or 0 < R < 1), the effective stress paths at large shear
stant confining pressure. VP200 is used to represent a test
strains are bounded by the peak state line (PSL) and the
in which the specimen undergoes vertical compression with
critical state line (CSL). These observations are consistent
r2 ¼ 200 kPa. Table 2 shows the test matrix in this study.
with those of Guo and Su (2007), Wan et al. (2007) and
Wan and Guo (2004) based on laboratory test results and
4. Effective stress paths and stress-strain responses constitutive simulations. It should be noted that the ulti-
mate case with # ! 1 is the isotropic compression. The
Fig. 2 presents the stress paths in the s-t space along a stress-strain responses along different dilative strain paths
variety of imposed strain paths ranging from extreme dila- are presented in Fig. 3(b).
tion (# ¼ 0:40; R ¼ 0:43) to extreme contraction (# ¼ 21; The deformation states corresponding to the two types
R ¼ 1:1). The initial consolidation pressure was 170 kPa. of contractant strain paths, the VEC and VCC tests, are
The critical state line (CSL) and the failure envelope (PSL) illustrated in Fig. 4. On the s-t plane, the strain paths in
corresponding to the peak state on the stress-strain curves the VEC and VCC tests are separated by the stress path
in drained compression tests under a constant confining of K0-compression at # ¼ 1 and R ! 1, as shown in
pressure are also presented in Fig. 2. The peak and critical Fig. 2. The stress-strain responses, presented in Fig. 3(c)

state friction angles are determined as up ¼ 23:6 and and (d) for VEC and VCC tests, are very distinctive

ucv ¼ 16 , respectively. The stress-strain responses in differ- depending on the deformation. For VEC tests with # > 0
ent tests are presented in Fig. 3. and R > 1, a peak value of t=s is obtained, with the ulti-
One immediately observes that the material response is mate deformation state being different from the critical
not bounded by the stress paths in the constant lateral con- state. As the value of R increases, the lateral strain tends
fining pressure (akin to a drained test) and the isochoric to decrease and gradually approaches the (biaxial) K0-
(akin to an undrained test of a saturated specimen) condi- condition with e2 ¼ 0. For strain paths for VCC with
tions. As the imposed strain ratio changes, the material # > 0 and R < 1, the t/s ratio increases monotonically

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
6 J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx

0.4 0.5

0.4
0.3
0.3
σ2=400kPa
0.2 σ2=300kPa

t/s
0.2 ℜ=54
t/s
σ2=200kPa
ℜ=10.8
0.1 0.1
e0=0.16 ℜ=1.0
e0=0.183
0 0
0 0.05 0.1 0.15 0.2 0.25 0 0.03 0.06 0.09 0.12 0.15
γ γ
(a) (b)
0.5
0.5
0.4 0.4
0.3 0.3
t/s

t/s
0.2 ℜ=−5.4 0.2 ℜ=1.0
ℜ=−2.2 ℜ=0.76
0.1 ℜ=−1.1 0.1 ℜ=0.43
ℜ=1.0 e0=0.183
e0=0.183
0 0
0 0.05 0.1 0.15 0 0.03 0.06 0.09 0.12 0.15
γ γ
(c) (d)
Fig. 3. Stress-strain responses in (a) compression tests with r2 = constant; (b) VED tests with 0 < R 6 1; (c) VEC tests with R P 1 and (d) VCC tests
with R < 1.

in different contact networks when shearing takes place


along various strain paths.

5.1. Internal structure evolution of overall contact network

5.1.1. Fabric evolution in VED and VEI tests


The degree of fabric anisotropy of the overall contact
network is characterized by q/ ¼ /1  /2 , in which /1
and /2 are the principal components of the fabric tensor
in the overall contact network. The directions of /1 and
(a) (b) /2 are coaxial with those of r1 and r2 , respectively. As
shown in Fig. 5(a), the value of q/ increases with the shear
Fig. 4. Two types of forced contractant tests: (a) the VEC test, and (b) the
VCC test. strain initially until it approaches the peak value. There-
after, q/ deceases as shearing continues and approaches a
constant at large strain. This trend is similar to the evolu-
with shear strain, gradually approaching a constant that tion of the deviatoric stress. However, both the peak value
depends on the strain ratio R and is smaller than sin ucv . and the ultimate value of q/ at large shear strains depend
These observations are generally consistent with the find-
on the imposed strain ratio R. In particular, the peak
ings of Wan et al. (2007).
and ultimate values of q/ increase as R decreases, implying
that the contact orientations show most intensive direc-
tional preference to resist deformation in the VED tests
5. Fabric evolution (0 < R < 1 or # < 0), in which the specimen collapses
much easier than in other tests. The evolution of q/ with
As the mechanical behavior of granular materials is a the stress ratio (t/s) is plotted in Fig. 5(b), in which the
macroscopic summation of all particle level activities, it is deformation starts from an initially isotropic stress state.
affected by the evolution of the spatial arrangement of par- Prior to the peak stress ratio, q/ increases with the increase
ticles and the distribution of inter-particle contact orienta- of (t/s) and the overall contact network resists deformation
tions. This section discusses the evolution of fabric tensors by rearranging contact orientation with more contacts in

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx 7

0.15 0.15
ℜ=1.0
ℜ=0.92
0.1 0.1 ℜ=0.86
ℜ=0.76
ϕ1 – ϕ2

ϕ1 – ϕ 2
ℜ=0.65
0.05 ℜ=1.0 ℜ=0.92 0.05 ℜ=0.43
ℜ=0.86 ℜ=0.76 e0=0.183
e0=0.18
0 0
0 0.05 0.1 0.15 0 0.1 0.2 0.3 0.4 0.5
γ t/s
(a) (b)
Fig. 5. Evolution of fabric anisotropy in the overall contact network (a) with shear strain, and (b) with stress ratio in the VED and VEI tests.

0.15 0.15
VED ℜ=10.8
0<ℜ<1 ℜ=2.2
VEI ℜ=1.0
0.1 ℜ=1 0.1 ℜ=0.86
ϕ1– ϕ2

ϕ 1 – ϕ2
ℜ=0.76
VEC ℜ=0.65
0.05 ℜ>1 0.05 ℜ=0.43

VCC
ℜ<-1 e0=0.183
0 0
0 0.2 0.4 0.6 0 0.1 0.2 0.3 0.4 0.5
t/s t/s
(a) (b)
Fig. 6. (a) Evolution of the fabric deviator for the overall contact network with t/s; and (b) the best fit of pre-peak data in different tests.

the vertical direction. At the peak stress state, the overall vide stronger support to the strong force chains in the ver-
contact network reaches its maximum bearing capacity tical direction. When the value of R is large enough, the
and buckles. After the peak, a reduction of q/ is observed weak force chains in the lateral direction will always be
as t/s decreases and gradually approaches the critical state. stable during deformation and provide sufficient support
Before the peak stress ratio, the relation between q/ and t/s to the strong force chains without collapsing. Correspond-
is non-linear, which is consistent with the conclusion of ingly, the stress paths could not approach the peak state
Mehrabadi and Nemat-Nasser (1983). Nevertheless, after line, for example, in the test with strain path of R ! þ1
the peak stress ratio, the variation of q/ with t/s ratio can in Fig. 2.
be described as a linear expression approximately, as
shown in Fig. 5. 5.1.3. Fabric evolution in VCC tests
In the VCC tests, both e_ 1 and e_ 2 are positive (compres-
sion) with e_ 1 > e_ 2 > 0. When compared with other tests,
5.1.2. Fabric evolution in VEC tests the enhanced lateral constraint in the VCC tests helps to
As shown in Fig. 6(a), significant fabric anisotropy build more contacts and increase the contact forces in the
develops in VED tests since a rapid lateral extension causes lateral direction. As a result, the directional variation of
a quick reduction of lateral constraint, which in turn the contact distribution decreases, which results in a
induces more contacts to align in the vertical direction to reduced magnitude of fabric anisotropy in the overall con-
resist deformation. Thereafter, failure of the overall con- tact network; as can be observed in Fig. 6(a).
tact network happens after the fabric anisotropy
approaches its maximum. The lower fabric anisotropy in
the VEC tests than that in the VED and VEI tests is due 5.1.4. Effect of stress on fabric evolution in overall contact
to the smaller extension rate (_e2 ¼ _e1 =R; R > 1) in the network
lateral direction than the vertical compression rate e_ 1 . With Fig. 6(b) presents the evolution of fabric deviator
the increase of R, the horizontal extension rate j_e2 j becomes q/ ¼ /1  /2 with respect to the stress ratio t/s in the
smaller, which tends to produce relatively stable force VED, VEI test and some VEC tests. A regression analysis
chains in the direction of r2. In other words, with an is performed to obtain a quantitative description for the
increase of R, the force chains in the lateral direction pro- fabric evolution. Prior to the peak stress state, the variation

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
8 J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx

of q/ with t/s and the imposed strain ratio can be described 5.2. Evolution of strong sub-network fabric
by
n 5.2.1. Fabric evolution in VED and VEI tests
/1  /2 ¼ Aðt=sÞ ð4Þ
Fig. 8 presents the evolution of fabric deviator
where A and n are functions of the imposed strain ratio R qs/ ¼ /s1  /s2 of the strong sub-network under different
or #. As shown in Fig. 7 (a), A and n can be reasonably conditions. Fig. 8(a) shows the variation of qs/ with the
expressed as A ¼ 1:5968  0:9446# and t/s ratio in the VED and VEI tests with R 2 ½0:43; 1:0 .
n ¼ 0:5049  0:7595#, respectively. Eq. (4) plays the role One remarkable finding is that a unique relation between
of the fabric evolution law that is important for a stress- qs/ and t/s can be defined prior to the peak deviator stress
strain model based on continuum mechanics.
ratio along all strain paths, which is different from the fab-
In the post-peak deformation stage, the value of q/
ric evolution of the overall contact network presented in
reduces with the decrease of t/s following a linear relation Fig. 5(b). The maximum induced anisotropy qs/;peak appears
q/ ¼ qmax
/ þ jp ½t=s  ðt=sÞmax with qmax
/ ¼ Aðt=sÞnmax ; as
simultaneously with (t/s)max. The post-peak data are
shown in Fig. 5(b). jp can be expressed as a linear function slightly scattered; however, they localize in a narrow band
of the imposed strain ratio jp ¼ 0:5054# þ 0:2377; see that almost overlaps the pre-peak data. Since /s1 þ /s2 ¼ 1,
Fig. 7(b). The reason that Eq. (4) does not hold true for the following relation can be used to describe the variation
the post-peak deformation stage may be attributed to the of qs/ with the t/s ratio:
different patterns of fabric evolution in the strong and weak
sub-networks as well as the potential deformation instabil- /s1  /s2 t
s ¼ j ð5Þ
ity. Owing to the rigid boundary conditions adopted in the /1 þ /2
s
s
DEM simulations, the post-peak deformation is still
homogeneous without noticeable localized or diffused where j is constant. The best estimate of j ¼ 1 can be
deformation patterns, and the deformation instability man- applied to different strain paths, as shown in Fig. 8(a).
ifests itself as the collapse or buckling of force chain Eq. (5) is verified by tests along imposed strain path
segments. R ¼ 0:7 for different initial void ratios in Fig. 8(b) and

0.3 3
A A = -0.7595ϑ + 0.5049
n
0.2 2
n
κp

0.1 n = -0.9446ϑ + 1.5968


A,

1
κ p = 0.5054ϑ+ 0.2377

0 0
-0.3 -0.2 -0.1 0 0.1 -0.6 -0.3 0 0.3 0.6
ϑ = dεv /dγ -ϑ = -dεv /dγ
(a) (b)
Fig. 7. Best fits of (a) the parameters A, n; and (b) the parameter of jp.

0.5 0.4 0.4


ℜ∈[0.43,1.0] e0=0.16, 0.19, 0.22, 0.25, 0.28 σ2=200kPa, 300kPa,400kPa
0.4 0.3
0.3
ϕs1 – ϕs2

ϕ s1 – ϕ s 2

0.3
ϕ s1 – ϕ s2

0.2 0.2
0.2
0.1 0.1
0.1 e0=0.16
ℜ= 0.7
0 0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4
t/s t/s t/s
(a) (b) (c)

Fig. 8. Evolution of the fabric deviator in the strong sub-network with stress ratio t/s (a) in the VED and VEI tests (# 6 0), (b) at different initial
porosities, and (c) under different confining pressures.

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx 9

by conventional biaxial compression tests with constant r2 VED tests, a peak and a residual anisotropy are
(VP200, VP300 and VP400) on specimens with an initial observed along strain paths with relatively small R val-
void ratio of 0.16 in Fig. 8(c). In both cases, Eq. (5) can ues, for example R ¼ 2:2. Eq. (5) can be used to describe
be used to describe the relation between t=s and the prior peak qs/
ðt=sÞ relation. Moreover, the occur-
ð/s1  /s2 Þ, regardless of the initial porosity of the specimens rence of (t/s)max and qs/;max implies that the maximum
and the variation of the dilatancy rate dev =dc in the con- resistance to deformation of the strong sub-network is
ventional biaxial compression tests. With j ¼ 1 holding mobilized due to intensive reduction of the lateral
true for all cases, Eq. (5) can be alternatively expressed as constraints.
/s1 : /s2 ¼ r1 : r2 . However, when R > 54, the expansion of the specimen
Eq. (5) can also be used to describe the variation of the in the lateral direction is relatively small. As a result, the
strong sub-network anisotropy in the post-peak deforma- strong force chain network can resist deformation without
tion stage, as shown in Fig. 8a for tests along controlled approaching its maximum capacity due to the strong sup-
strain paths and Fig. 8b for tests with constant confining port of the horizontal weak force chains. Therefore, qs/
stresses. The variation of qs/ with the shear strain depends increases monotonically with the stress ratio (t/s) following
on the rate of volume change, while a critical state for Eq. (5). Compared with the scattered q/
t=s relations for
qs/;cr is identified at large shear strain level in Fig. 8. The the overall contact network in Fig. 6, Fig. 9(a) suggests a
decrease of qs/ implies a degradation, even collapse of strong influence of t/s on the internal structure of strong
strong sub-network, which results in a decrease of shear sub-network in tests along contractant strain paths with
resistance and potential deformation instability. The corre- R > 1. In addition, the ultimate fabric deviator qs/;ult
lation between the evolution of strong sub-network aniso- depends on the value of R in these cases. qs/;ult is higher
tropy and deformation instability will be discussed in later than the critical value qs/;cr in the VED and VEI tests.
sections.
5.2.3. Fabric evolution in VCC tests
5.2.2. Fabric evolution in VEC tests For imposed contractant strain paths with # > 0 and
In VEC tests, shearing takes place along contractant R < 1 in the VCC tests, qs/ increases monotonically with
strain paths, the qs/
ðt=sÞ curves are all in a narrow t/s and the shear strain; as shown in Figs. 9(b) and (d). The
band around the line of /s1  /s2 ¼ t=s; as shown in ultimate fabric deviator qs/;ult increases as the strain paths
Fig. 9(a). Similar to the development of anisotropy in moves toward K0-compression at jRj ! 1. As the jRj

0.5
0.5 ℜ=−1.1
e0=0.183
0.4 ℜ=−2.2
0.4
ℜ=−2.7
0.3 0.3 ℜ=−3.2
ϕ s1 – ϕs 2

ϕs1– ϕs2

ℜ=−5.4
0.2 0.2
ℜ∈[2.2,+∞]
0.1 0.1
ϑ∈[0.38,1.0]
e0=0.183
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
t/s t/s
(a) (b)
0.5 0.5

0.4 0.4
ℜ=−1.1
ℜ=2.2 0.3 ℜ=−2.2
0.3
ϕ s 1– ϕ s 2

ℜ=10.8 ℜ=−2.7
ϕ s 1 – ϕs 2

0.2 ℜ=21.6 0.2 ℜ=−3.2


ℜ=32.4 ℜ=−5.4
0.1 0.1
ℜ=54.0 e0=0.183
e0=0.183 0
0
0 0.02 0.04 0.06 0.08
0 0.05 0.1 0.15
γ γ
(c) (d)
Fig. 9. Evolution of qs/ ¼ /s1  /s2 with t/s in VEC tests (a), in VCC tests (b), with the shear strain in VEC tests (c) and in VCC tests (d).

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
10 J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx

ℜ =1.0 K0 -1
0.6 ψmax

-0.5
0.4
ϕs1– ϕs2
0

(%)
VEC VCC upper limit σ2=400kPa
0.2 σ2=400KPa σ2=300kPa
0.5

εv
σ2=300KPa σ2=200kPa
σ2=200KPa
e0=0.16
0 1
-0.5 0 0.5 1 1.5 2 2.5 3 0 0.05 0.1 0.15 0.2 0.25
dεv/dγ γ
(a) (b)

Fig. 10. (a) The upper limits of qs/ ¼ /s1  /s2 ; and (b) Volumetric strain in biaxial compression tests when r2 = constant.

value increases, the volume contraction becomes dominant provides an adequate measure of the geometrical aniso-
and the effect of shearing decreases. As a special case when tropy (Radjaı̈ et al., 1998):
R ¼ 1, the specimen is subjected to hydrostatic compres- 1  
sion. One expects smaller qs/;ult values as R increases, which Es ðnÞ  1 þ Asc cos 2ðh  hsc Þ ð7Þ
2p
is confirmed by the results in Figs. 9(b) and (d). With the
decrease in the R value, the lateral constraint becomes where Asc and hsc define the amplitude of anisotropy and the
weaker, and hence severe anisotropy develops in the strong major principal direction of the fabric tensor for the strong
sub-network. sub-network. It can be shown that /s1  /s2 ¼ Asc =2. The
non-negative requirement of Es ðnÞ requires Asc 6 1, which
yields qs/;max 6 0:5 and is consistent with the results pre-
5.2.4. Maximum anisotropy of strong sub-network sented in Fig. 10(a).
Fig. 10 show the variation in the maximum fabric ani-
sotropy qs/;max in the strong sub-network with # ¼ dev =dc 5.3. Evolution of weak network fabric
in controlled strain path tests. The data can be divided
into two sections by the K0-compression (R ! 1) The evolution qw/ ¼ /w1  /w2 of the weak sub-network
where the value of qs/;max approaches its maximum value. with the t/s ratio and shear strain along different strain
The right section of the data points corresponds to the paths are presented in Fig. 11. Minor anisotropy in the
VCC tests (R < 1) in which the specimen has compres- weak sub-network develops in VED tests (0 < R < 1), with
sive deformation in the directions of both r1 and r2 . The the major principal direction of /wij perpendicular to the
qs/;max tends to decrease as # ¼ dev =dc increases in these direction of r1 and hence the direction of /s1 in the strong
tests, which is induced by the increased fraction of strong sub-network. Following an initial buildup of anisotropy
contacts in the lateral direction due to the over- upon loading, qw/ reaches a maximum value and then
constraint imposed by deformation in the direction of decreases with continuous deformation. The initial buildup
r2 . However, qs/;max increases with the increase of # in of qw/ along dilatant strain paths can be described as a lin-
the left section of the data points in VEC tests, in which ear function of t/s in the form of:
the specimen is allowed to expand in the direction of r2 .
In biaxial tests with constant r2 , the induced anisotropy /w1  /w2 t
¼ jw ð8Þ
is close to that along dilative strain paths with /w1 þ /w2 s
# ¼  sin wmax in which wmax is the maximum dilatancy
in which jw is a coefficient that varies with the value of R.
angle defined as sin wmax ¼ dev =dc in conventional biax-
As shown in Fig. 12(a), jw ranges from 0.3 to 0.45 when the
ial tests (see Fig. 10b).
dilation rate increases form # ¼ 0:4 (R ¼ 0:43) to # ¼ 0
It can be shown that the theoretical maximum value of
(R ¼ 1:0). After the peak anisotropy, the degree of aniso-
fabric deviator in the strong sub-network is 0.5. For contin-
tropy diminishes gradually and the weak sub-network
uous distribution of contact normal, /sij defined in Eq. (2)
eventually becomes isotropic at large strain in tests of high
can be alternatively expressed as
dilation rate (e.g., R ¼ 0:43–0:65); as shown in Fig. 12(a).
Z 2p For the VEC tests (R > 1) and VCC tests (R < 1), the
/ij ¼
s
Es ðnÞni nj dh ð6Þ induced anisotropy in the weak sub-network tends to
0
increase significantly as the deviatoric stress ratio t/s
with Es ðnÞ being the probability density function defining increases. With the increase of shear strain, the value of
the directional distribution of contact normal in the strong qw/ approaches a limit depending on the imposed strain
sub-network. The first-order Fourier expansion of Es ðnÞ ratio R; see Fig. 12(a).

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx 11

0.4 0.4
ℜ=-2.2 ℜ=-5.4 ℜ=-2.2
ℜ=-5.4
0.3 ℜ=10.8 0.3 ℜ=10.8
ℜ=2.2
ℜ=2.2 ℜ=1.0
qwϕ

qwϕ
0.2 0.2
ℜ=0.76
ℜ=1.0 ℜ=0.43
0.1 ℜ=-1.1 0.1 ℜ=-1.1
ℜ=0.76
ℜ=0.43
0 0
0 0.05 0.1 0.15 0 0.15 0.3 0.45
γ=ε1-ε2 t/s
(a) (b)
Fig. 11. Evolution of qw/ ¼ /w1  /w2 (a) with shear strain, and (b) with stress ratio along various strain paths.

ℜ=0.43 ℜ=1.0 ℜ=10.8 ℜ=−5.4


0.2
ℜ=1.0
ℜ=0.92
0.15 ℜ=0.86
W
ℜ=0.76
0.1
qwϕ

S
0.05

0 O
0 0.1 0.2 0.3 0.4 0.5
t/s
(a) (b)
Fig. 12. (a) Linear relation between qw/ and t/s prior to the qw/;max in VED and VEI tests; and (b) the histogram of the strong sub-network (S), the weak sub-
network (W) and the overall contact network (O) when c = 3%.

Different trends are observed regarding the evolution of in Fig. 6. However, due to the non-uniform distribution
qw/ with the t/s ratio in the VEC and VCC tests. As shown in of contact forces, a strong and a weak sub-network can still
Fig. 12(b), for the VEC tests, the degree of fabric aniso- be identified with a certain level of anisotropy, which is
tropy within the weak sub-network maintains at a steady stronger for the strong sub-network and weaker for the
level in the post-peak t/s-ratio stage. The level of aniso- weak sub-network in the r1 direction. The minor aniso-
tropy is much lower than that developed in the strong tropy of the overall contact network requires that the fabric
sub-network. On the other hand, when compression takes anisotropies in the strong and weak networks are both
place in both directions with e_ 1 > e_ 2 > 0 and R < 1 small, as demonstrated in Fig. 13.
(i.e., VCC tests), the induced anisotropy in the weak sub-
network reaches a much higher level as the t/s increases if 6. Fabric evolution and deformation instability
the strain paths are close to K 0 -compression. As a special
case, no significant anisotropy is observed in the weak sub- Strain softening and instability of geomaterials may lead
network when a specimen is subjected to isotropic com- to catastrophic deformation, including failure. In mechan-
pression (R ¼ 1). It should be emphasized that the major ical terms, stability is obtained if a small stress increment
principal direction of the weak sub-network is perpendicu- yields a small strain increment, while instability is defined
lar to that of the strong sub-network. as a behaviour in which large plastic strains develop rapidly
Fig. 13 compares the evolution of fabric in the strong due to the inability of material to sustain a given stress or
and weak sub-networks along contractant strain paths in load. The theoretical basis for the concept of stability is
which # > 0 and R < 1. For R ¼ 1.1 to 5.7, the evo- Drucker’s postulate of non-negative second-order plastic
lution of qs/ and qw/ are very close, especially at R ¼ 1:1. It work, i.e., d2 Wp ¼ drij depij > 0 (Drucker, 1957). Hill
should be noted that the strain path of R ¼ 1 corre- (1958) extended Drucker’s postulate to the second-order
sponds to isotropic compression. For an initially isotropic increment of total work. According to Hill (1958), for
(or weakly anisotropic) contact network, the induced ani- any (drij ; deij ) linked by their constitutive relation, a mate-
sotropy of the overall contact network when R ¼ 1:1 rial is stable when d2 W ¼ drij deij 6 0. Hence, a positive
and 2.2 is expected to be minimal, as is demonstrated second-order work constitutes a sufficient, but not neces-

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
12 J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx

0.6 0.5
qwK=-
ϕ ℜ=-5.4
1.1
0.4
qsK=1.
ϕ
0.4 0s
0.3
qwϕ, qsϕ ℜ=-2.2

q sϕ
ℜ=1.0 0.2 ℜ=1.0
0.2
ℜ=−1.1
ℜ=-1.1 0.1 ℜ=−2.2
ℜ=−5.4
0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
t/s qwϕ
(a) (b)
Fig. 13. (a) Evolution of qs/ and qw/ with t/s; and (b) Relation between qs/ and qw/ in VCC tests.

sary, condition of stability. When d2 W 6 0, the effective Fig. 14 presents the variation of ðr1  r2 =RÞ and its
collapse of a homogeneous sample depends on the loading relation with the fabric deviator in the strong and weak
direction and the controlled loading variables (Nicot et al., sub-networks. In particular, the variation of ðr1  r2 =RÞ
2013; Prunier et al., 2009). Alternatively, d2 W 6 0 can be with axial strain e1 along various strain paths is plotted
interpreted as a bifurcation criterion. At a bifurcation in Fig. 14(a), where the peak points corresponding to the
point, the response path can be unique or nonunique and onset of deformation instability according to Eq. (10) can
stable or unstable. For example, in the tests along propor- be easily identified in the VED tests (0 < R < 1). For the
tional strain paths, the failure corresponding to the vanish- VEC (R > 1) and VCC (R < 1) tests, however, the value
ing of the second-order work is thought to be non-effective of ðr1  r2 =RÞ tends to increase monotonically with e1 ,
(Nicot et al., 2013). indicating that deformation is always stable. Fig. 14(b) pre-
Attempts have been made to investigate material insta- sents the correlation between qs/ ¼ /s1  /s2 and
bility in granular assemblies on the microscopic scale in ðr1  r2 =RÞ in the VED tests with R ¼ 0:43–1. The thicker
terms of the microstructure of the overall contact network; dotted segment on each curve represents states of unstable
see, e.g., Wan et al. (2007), Nicot and Darve (2006, 2005) deformation. An immediate observation is that the strong
and Nicot et al. (2015). It has been revealed that the origin sub-network is responsible for deformation instability since
of potential instabilities as a macro-scale phenomenon may the peak value of qs/ is reached when d2W = 0. More specif-
be directly related to the constitutive nature of the local ically, for tests of high dilation rate with R ¼ 0:43–0:65, qs/
contact model at particle contacts (Nicot and Darve, increases initially with ðr1  r2 =RÞ, both reaching their
2006). In addition, Nicot et al. (2007) observe that changes maximum simultaneously; see Figs. 14(a) and (b). Refer-
in the fabric of the medium, associated with rearrange- ring to Figs. 2 and 3(b), for high dilative strain paths with
ments caused by sliding and rolling, cause nonequivalence R ¼ 0:43–0:65, d2W = 0 occurs at the ultimate deforma-
between the microscopic and macroscopic second-order tion state when t=s ¼ ðt=sÞmax and the shear resistance
works. The following analyses, however, will focus on the approaches zero. For lower rates of dilation with
variation of fabric in different contact networks when the R ¼ 0:76–0:86, qs/ and ðr1  r2 =RÞ also reach their maxi-
macroscopic second-order work vanishes. mum at the same time. However, the degradation of the
Following the stability analysis for sand tested along strong sub-network with the reduced degree of anisotropy
proportional strain paths undertaken by Wan et al. takes place following the onset of deformation instability;
(2007), Darve and Laouafa (2000) and Wan et al. (2005), as shown in Fig. 14(b). In the CEI test with R ¼ 1, the peak
the second-order work under biaxial stress conditions can value of qs/ develops during deformation that is always
be expressed as
  stable according to Fig. 14(b) since ðr1  r2 =RÞ increases
dr2  r2 monotonically. These observations indicate that the onset
d2 W ¼ dr1  de1 ¼ d r1  de1 ð9Þ
R R of deformation instability is accompanied by a degradation
or progressive collapse of the strong sub-network. How-
with R ¼ de1 =de2 as defined previously. With the strain
ever, the degradation of the strong sub-network does not
increment de1 being non-negative, the condition d2 W ¼ 0 guarantee unstable deformation. In other words, the degra-
is satisfied when dðr1  r2 =RÞ ¼ 0, for which dation of strong sub-network is a necessary but not suffi-
ðr1  r2 =RÞ ¼ max ð10Þ cient condition for deformation instability.
Fig. 14(c) displays the evolution of qw/ ¼ /w1  /w2 with
The above equation implies that instability starts at the
peak point of the curve obtained by plotting ðr1  r2 =RÞ ðr1  r2 =RÞ. For dilation rates in the range of
versus axial strain e1 . R ¼ 0:43–0:86, the highest value of qw/ is observed before

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx 13

ℜ=−1.1 ℜ=10.8
ℜ=2.2
1100
ϑ>0

σ1 – σ2 / ℜ, kPa
600 d2W>0
ℜ=1
ℜ=0.76
100
ϑ <0
ℜ=0.43 ℜ=0.65 d2W=0
-400
0 0.02 0.04 0.06 0.08 0.1
Axial strain

(a)
0.5 0.2
ℜ=0.76 ℜ=1
0.4 ℜ=0.65 ℜ=1
0.15
ℜ=0.86
0.3 ℜ=0.43
ℜ=0.76
q sϕ

ℜ=0.86 0.1 ℜ=0.65

qwϕ
0.2
ℜ=0.43
0.1 0.05

0
0
-400 -200 0 200 400 600
-400 -200 0 200 400 600
(σ1-σ2/ℜ), kPa
(σ1-σ2/ℜ), kPa
(b) (c)
Fig. 14. (a) Variation of r1  r2 =R with axial strain, (b) evolution of fabric deviator for the strong sub-network and (c) for the weak sub-network with
r1  r2 =R in VED tests (the thick dotted line represents the d2W = 0).

0.5 0.5

0.4 0.4
ℜ=2.2 0.3 ℜ=−1.1
0.3
ℜ=10.8 ℜ=−2.2
q sϕ

q sϕ

0.2 ℜ=21.6 0.2 ℜ=−2.7


ℜ=32.4 ℜ=−3.2
0.1 ℜ=54 0.1 ℜ=−5.4

0 0
0 1000 2000 3000 4000 0 2000 4000 6000
(σ1-σ2/ℜ), kPa (σ1-σ2/ℜ), kPa
(a) (b)
0.4 0.4

0.3 0.3
ℜ=−1.1
ℜ=2.2 ℜ=−2.2
ℜ=10.8
qwϕ

0.2 0.2 ℜ=−2.7


ℜ=21.6
qwϕ

ℜ=−3.2
ℜ=32.4
0.1 0.1 ℜ=−5.4
ℜ=54.0

0 0
0 1000 2000 3000 4000 0 2000 4000 6000
(σ1-σ2/ℜ), kPa (σ1-σ2/ℜ), kPa
(c) (d)
Fig. 15. Evolution of qs/ ¼ /s1  /s2 with r1  r2 =R in VEC tests (a) and VCC tests (b), and evolution of qw/ ¼ /w1  /w2 with r1  r2 =R in VEC tests (c)
and VCC tests (d).

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
14 J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx

ðr1  r2 =RÞmax is achieved. In all these cases, a significant (3) The strong sub-network provides resistance to devia-
decrease in qw/ takes place when deformation becomes tor stresses of a granular assembly, while the weak
unstable. In other words, the degradation of the weak force chain network and its evolution are significantly
sub-network is initiated before deformation becomes affected by the imposed dilatancy rate.
unstable. In the case of the VEI test (R ¼ 1), deformation (4) The instability of a granular assembly is closely
is always stable and practically no degradation of qw/ related to the degradation or even collapse of the
occurs. weak force chain network, even though the strength
Fig. 15 compares the relation between ðr1  r2 =RÞ and of the assembly depends on the strong sub-network.
fabric deviators in the weak and strong sub-networks in the
VEC (R > 1) and VCC tests (R < 1). In all cases, the
deformation is always stable since ðr1  r2 =RÞ never Acknowledgement
reaches the maximum in these tests. qw/ increases monoton-
Funding provided by the Natural Sciences and Engi-
ically and approaches a steady state that depends on the
neering Research Council of Canada and the China Schol-
ultimate stress states; as shown Figs. 15(c) and (d). On
arship Council (CSC) is gratefully acknowledged.
the other hand, qs/ only increases monotonically in the
VCC tests; see Fig. 15(b). However, in the VEC tests, a
References
peak value of qs/ develops at R ¼ 2:2 followed by a
decrease in qs/ ; as shown in Figs. 15(a) and 9(c). Antony, S.J., Momoh, R.O., Kuhn, M.R., 2004. Micromechanical
Based on the discussion about the results presented in modelling of oval particulates subjected to bi-axial compression.
Figs. 14 and 15, one may conclude that the degradation Comput. Mater. Sci. 29 (4), 494–498. https://doi.org/10.1016/
j.commatsci.2003.12.007.
of the strong sub-network (i.e., the decrease of qs/ ) is a nec- Bathurst, R.J., Rothenburg, L., 1988. Micromechanical aspects of
essary but not a sufficient condition of deformation insta- isotropic granular assemblies with linear contact interactions. J. Appl.
bility. In other words, deformation instability is Mech. 55 (1), 17–23. https://doi.org/10.1115/1.3173626.
accompanied by strong network degradation but not the Chang, C.S., Chao, S.J., Chang, Y., 1995. Estimates of elastic moduli for
granular material with anisotropic random packing structure. Int. J.
opposite. On the other hand, a decrease in qw/ is the indica- Solids Struct. 32 (14), 1989–2008. https://doi.org/10.1016/0020-7683
tor of deformation instability in tests along all imposed (94)00225-L.
strain paths. Christoffersen, J., Mehrabadi, M.M., Nemat-Nasser, S., 1981. A microme-
chanical description of granular material behavior. J. Appl. Mech.
Trans. ASME 48 (2), 339–344. https://doi.org/10.1115/1.3157619.
7. Concluding remarks
Cundall, P.A., Strack, O.D.L., 1979. A discrete numerical model for
granular assemblies. Géotechnique 29 (1), 47–65. https://doi.org/
In this study, the behaviour of granular materials in 10.1680/geot.1979.29.1.47.
biaxial tests along various strain paths was simulated using Darve, F., Laouafa, F., 2000. Instabilities in granular materials and
the discrete element method. Without attempting to repro- application to landslides. Mech. Cohesive-Frictional Mater. 5 (8), 627–
652. https://doi.org/10.1002/1099-1484(200011)5:8<627::AID-
duce any laboratory test results, this study sheds light on
CFM109>3.0.CO;2-F.
the capacity of discrete element methods to simulate the Drucker, D.C., 1957. A definition of stable inelastic material. J. Appl.
constitutive features observed during laboratory tests, with Mech. 26 (1), 101–106. https://doi.org/10.2307/1146143.
the microstructural context revealing information about Fu, P., Dafalias, Y.F., 2015. Relationship between void- and contact
both the deformation instability and failure. To character- normal-based fabric tensors for 2D idealized granular materials. Int. J.
Solids Struct. 63, 68–81. https://doi.org/10.1016/j.ijsolstr.2015.02.041.
ize the microstructure and its evolution in granular assem-
Fu, P., Dafalias, Y.F., 2011. Fabric evolution within shear bands of
blies, the overall contact network was split into a strong granular materials and its relation to critical state theory. Int. J.
and a weak sub-network, both described by two indepen- Numer. Anal. Methods Geomech. 35 (18), 1918–1948. https://doi.org/
dent fabric tensors. The correlation between the evolution 10.1002/nag.988.
of different contact networks and material behaviour Gao, Z., Zhao, J., Li, X.S., Dafalias, Y.F., 2014. A critical state sand
plasticity model accounting for fabric evolution. Int. J. Numer. Anal.
including deformation instability was examined. The fol-
Methods Geomech. 38 (4), 370–390. https://doi.org/10.1002/nag.2211.
lowing conclusions are obtained: Guo, P., Su, X., 2007. Effect of dilatancy on instability, pre-instability
strain softening of sand along proportional strain paths. Soils Found.
(1) The fabric evolution of the overall contact network 47 (4), 757–770.
along an imposed strain path can be related to the Guo, P.J., Stolle, D.F.E., 2005. On the failure of granular materials with
fabric effects. Soils Found. 45 (4), 1–12.
stress ratio (t/s). However, such relation varies with
Hill, R., 1958. A general theory of uniqueness and stability in elastic-
the imposed dilatancy rate dev =dc . plastic solids. J. Mech. Phys. Solids 6 (3), 236–249. https://doi.org/
(2) Along dilatant strain paths, a unique fabric-stress 10.1016/0022-5096(58)90029-2.
relation r1 =r2 ¼ /s1 =/s2 , or equivalently Huang, X., Hanley, K.J., O’Sullivan, C., Kwok, C.Y., 2014. Exploring the
ð/1  /2 Þ=ð/s1 þ /s2 Þ ¼ t=s, is established in the
s s influence of interparticle friction on critical state behaviour using
DEM. Int. J. Numer. Anal. Methods Geomech. 38 (12), 1276–1297.
strong sub-network. The maximum value of qs/ for https://doi.org/10.1002/nag.2259.
all tests occurs in K 0 -compression tests.

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001
J. Shi, P. Guo / Soils and Foundations xxx (2018) xxx–xxx 15

Jiang, M., Yu, H., Harris, D., 2006. Kinematic variables bridging discrete Oda, M., Konishi, J., Nemat-Nasser, S., 1982. Experimental microme-
and continuum granular mechanics. Mech. Res. Commun. 33 (5), 651– chanical evaluation of strength of granular materials: effects of particle
666. https://doi.org/10.1016/j.mechrescom.2005.06.013. rolling. Mech. Mater. 1 (4), 269–283. https://doi.org/10.1016/0167-
Kanatani, K.I., 1984. Stereological determination of structural aniso- 6636(82)90027-8.
tropy. Int. J. Eng. Sci. 22 (5), 531–546. https://doi.org/10.1016/0020- Prunier, F., Laouafa, F., Lignon, S., Darve, F., 2009. Bifurcation
7225(84)90055-7. modeling in geomaterials: from the second-order work criterion to
Kruyt, N.P., 2012. Micromechanical study of fabric evolution in quasi- spectral analyses. Int. J. Numer. Anal. Methods Geomech. 33 (9),
static deformation of granular materials. Mech. Mater. 44, 120–129. 1169–1202. https://doi.org/10.1002/nag.762.
https://doi.org/10.1016/j.mechmat.2011.07.008. Radjaı̈, F., Delenne, J.Y., Azéma, E., Roux, S., 2012. Fabric evolution
Kuhn, M.R., Sun, W.C., Wang, Q., 2015. Stress-induced anisotropy in and accessible geometrical states in granular materials. Granul. Matter
granular materials: fabric, stiffness, and permeability. Acta Geotech. 14 (2), 259–264. https://doi.org/10.1007/s10035-012-0321-8.
10 (4), 399–419. https://doi.org/10.1007/s11440-015-0397-5. Radjaı̈, F., Jean, M., Moreau, J.-J., Roux, S., 1996. Force distributions in
Lade, P.V., 2008. Failure criterion for cross-anisotropic soils. J. Geotech. dense two-dimensional granular systems. Phys. Rev. Lett. 77 (2), 274–
Geoenviron. Eng. 134 (1), 117–124. https://doi.org/10.1061/(ASCE) 277. https://doi.org/10.1103/PhysRevLett. 77.274.
1090-0241(2008) 134:1(117). Radjaı̈, F., Roux, S., Moreau, J.J., 1999. Contact forces in a granular
Li, X.S., Dafalias, Y.F., 2011. Anisotropic critical state theory: role of packing. Chaos An Interdiscip. J. Nonlinear Sci. 9 (3), 544–550.
fabric. J. Eng. Mech. 138 (3), 263–275. https://doi.org/10.1061/ https://doi.org/10.1063/1.166428.
(ASCE)EM.1943-7889.0000324. Radjaı̈, F., Wolf, D., Jean, M., Moreau, J.-J., 1998. Bimodal character of
Maeda, K., Hirabayashi, H., Ohmura, A., 2006. Micromechanical influence stress transmission in granular packings. Phys. Rev. Lett. 80 (1), 61–64.
of grain properties on deformation-failure behaviors of granular media https://doi.org/10.1103/PhysRevLett. 80.61.
by DEM. Geomech. Geotech. Particulate Media, 173–179. Rothenburg, L., Kruyt, N.P., 2004. Critical state and evolution of
Mehrabadi, M.M., Nemat-Nasser, S., 1983. Stress, dilatancy and fabric in coordination number in simulated granular materials. Int. J. Solids
granular materials. Mech. Mater. 2 (2), 155–161. https://doi.org/ Struct. 41 (21), 5763–5774. https://doi.org/10.1016/j.
10.1016/0167-6636(83)90034-0. ijsolstr.2004.06.001.
Mohamed, A., Gutierrez, M., 2010. Comprehensive study of the effects of Rothenburg, L., Selvadurai, A.P.S., 1981. A micromechanical definition
rolling resistance on the stress–strain and strain localization behavior of the Cauchy stress tensor for particulate media. Mech. Struct. media,
of granular materials. Granul. Matter 12 (5). https://doi.org/10.1007/ 469–486.
s10035-010-0211-x, 3527-3271. Satake, M., 1987. Graph-theoretical approach to the mechanics of
Muhunthan, B., Collins, I.F., 2003. On the relationship between stress– granular materials. In: Spencer, A.J.M. (Ed.), Continuum Models of
dilatancy, anisotropy, and plastic dissipation for granular materials. Discrete Systems. Taylor & Francis, Boston, pp. 163–172.
Géotechnique 53 (7), 611–618. https://doi.org/ Satake, M., 1978. Constitution of mechanics of granular materials
10.1680/geot.2003.53.7.611. through the graph theory. Contin. Mech. Stat. approaches Mech.
Nakai, T., 1989. An isotropic hardening elastoplastic model for sand Granul. Mater. 47–62.
considering the stress path dependency in three-dimensional stresses. Suiker, A.S.J., Chang, C.S., 2004. Modeling failure and deformation of an
Soils Found. 29 (1), 119–137. assembly of spheres with frictional contacts. J. Eng. Mech. 130 (3),
Ng, T., 2001. Fabric evolution of ellipsoidal arrays with different particle 283–293.
shapes. J. Eng. Mech. 127 (10), 994–999. https://doi.org/10.1061/ Thornton, C., 2000. Numerical simulations of deviatoric shear deforma-
(ASCE)0733-9399(2001) 127:10(994). tion of granular media. Géotechnique 50 (1), 43–53. https://doi.org/
Nguyen, N.S., Magoariec, H., Cambou, B., Danescu, A., 2009. Analysis 10.1680/geot.2000.50.1.43.
of structure and strain at the meso-scale in 2D granular materials. Int. Wan, R., Al-Mamun, M., Guo, P., 2007. Experimental investigation of
J. Solids Struct. 46 (17), 3257–3271. https://doi.org/10.1016/j. instabilities of granular materials in relation to dilatancy and fabric
ijsolstr.2009.04.019. issues. Bifurcations, Instabilities, Degradation in Geomechanics.
Nicot, F., Daouadji, A., Hadda, N., Jrad, M., Darve, F., 2013. Granular Springer, Berlin Heidelberg, pp. 71–93.
media failure along triaxial proportional strain paths. Eur. J. Environ. Wan, R., Pouragha, M., 2015. Fabric and connectivity as field descriptors
Civ. Eng. 17 (9), 777–790. https://doi.org/10.1080/ for deformations in granular media. Contin. Mech. Thermodyn. 27 (1–
19648189.2013.819301. 2), 243–259. https://doi.org/10.1007/s00161-014-0370-9.
Nicot, F., Darve, F., 2006. Micro-mechanical investigation of material Wan, R.G., Al-Mamun, M., Guo, P.J., 2005. How do fabric and dilatancy
instability in granular assemblies. Int. J. Solids Struct. 43 (11–12), affect the strength of granular materials, in: Proceedings of the
3569–3595. https://doi.org/10.1016/j.ijsolstr.2005.07.008. International Conference on Soil Mechanics and Geotechnical Engi-
Nicot, F., Darve, F., 2005. A multi-scale approach to granular materials. neering. pp. 863–868.
Mech. Mater. 37, 980–1006. https://doi.org/10.1016/j. Wan, R.G., Guo, P.J., 2004. Stress dilatancy and fabric dependencies on
mechmat.2004.11.002. sand behavior. J. Eng. Mech. 130 (6), 635–645. https://doi.org/
Nicot, F., Sibille, L., Donze, F., Darve, F., 2007. From microscopic to 10.1061/(ASCE)0733-9399(2004) 130:6(635).
macroscopic second-order work in granular assemblies. Mech. Mater. Zhao, J., Guo, N., 2013. Unique critical state characteristics in granular
39 (7), 664–684. https://doi.org/10.1016/j.mechmat.2006.10.003. media considering fabric anisotropy. Géotechnique 63 (8), 695–704.
Nicot, F., Sibille, L., Hicher, P.Y., 2015. Micro-macro analysis of granular https://doi.org/10.1680/geot.12.P.040.
material behavior along proportional strain paths. Contin. Mech. Zhu, H., Nicot, F., Darve, F., 2016. Meso-structure evolution in a 2D
Thermodyn. 27 (1–2), 173–193. https://doi.org/10.1007/s00161-014- granular material during biaxial loading. Granul. Matter 18 (1), 1–12.
0347-8. https://doi.org/10.1007/s10035-016-0608-2.
Oda, M., 1993. Inherent and induced anisotropy in plasticity theory of
granular soils. Mech. Mater. 16 (1–2), 35–45. https://doi.org/10.1016/
0167-6636(93)90025-M.

Please cite this article in press as: Shi, J., Guo, P., Induced fabric anisotropy of granular materials in biaxial tests along imposed strain paths, Soils Found.
(2018), https://doi.org/10.1016/j.sandf.2018.02.001

You might also like