You are on page 1of 10

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/226527995

Modeling of the simple shear deformation of sand: Effects of principal stress


rotation

Article  in  Acta Geotechnica · September 2009


DOI: 10.1007/s11440-009-0094-3

CITATIONS READS

25 1,200

3 authors:

Marte Gutierrez J. F. Wang


Colorado School of Mines City University of Hong Kong
304 PUBLICATIONS   4,622 CITATIONS    93 PUBLICATIONS   2,495 CITATIONS   

SEE PROFILE SEE PROFILE

Mitsutoshi Yoshimine
Tokyo Metropolitan University
33 PUBLICATIONS   1,696 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Proppant transport in a hydraulic fracture driven by low-viscous fluid View project

Hydro-mechanical response of deep tunnels in saturated ground View project

All content following this page was uploaded by Marte Gutierrez on 26 June 2015.

The user has requested enhancement of the downloaded file.


Acta Geotechnica (2009) 4:193–201
DOI 10.1007/s11440-009-0094-3

RESEARCH PAPER

Modeling of the simple shear deformation of sand:


effects of principal stress rotation
Marte Gutierrez Æ J. Wang Æ M. Yoshimine

Received: 10 November 2008 / Accepted: 7 May 2009 / Published online: 23 July 2009
Ó US Government 2009

Abstract The paper presents a simple constitutive model 1 Introduction


for the behavior of sands during monotonic simple shear
loading. The model is developed specifically to account for One of the most common types of in situ soil deformation
the effects of principal stress rotation on the simple shear is that of simple shear. Simple shear condition is pre-
response of sands. The main feature of the model is the dominant during earthquake shaking of level grounds when
incorporation of two important effects of principal stress on deformation propagates vertically upwards from the bed-
stress–strain response: anisotropy and non-coaxiality. In rock in the form of shear waves. Soils within localized
particular, an anisotropic failure criterion, cross-anisotropic failure zones also deform in simple shear. Simple shear
elasticity, and a plastic flow rule and a stress–dilatancy deformation can be simulated in element tests using the
relationship that incorporate the effects of non-coaxiality direct simple shear (DSS) device, or the hollow cylindrical
are adopted in the model. Simulations of published torsional (HCT) simple shear device. The main drawback
experimental results from direct simple shear and hollow of simple shear devices is the non-uniformity of the stress
cylindrical torsional simple shear tests on sands show the and strain distributions within the soil specimen. Despite
satisfactory performance of the model. It is envisioned that this shortcoming, simple shear testing continues to have
the model can be valuable in modeling in situ simple shear wide use in practice due to the relevance of the test to in
response of sands and in interpreting simple shear test situ conditions and the ease of carrying out the test.
results. Another drawback in the use of simple shear testing is
the difficulty of interpreting the test results. A key feature
Keywords Constitutive model  Dilatancy  of simple shear loading is that the principal stress direc-
Granular materials  Non-coaxiality  Simple shear tions are not fixed but rotate during shearing. As a result,
the orientations of the failure planes are not a priori known
and depend on the degree of principal stress rotation.
Simple shear devices, where the sample is constrained from
M. Gutierrez (&) deforming laterally, have only two degrees of freedom
Division of Engineering, Colorado School of Mines,
corresponding to the shear and vertical deformations. Thus,
Golden, CO 80401, USA
e-mail: mgutierr@mines.edu the principal stress rotation cannot be directly controlled
and only limited rotation can be achieved [22]. Experi-
J. Wang mental determination of the degree of principal stress
Department of Building and Construction,
rotation requires measurement of the variation of the lateral
City University of Hong Kong, Kowloon, Hong Kong
e-mail: jefwang@cityu.edu.hk confining stress during shearing. However, this measure-
ment is difficult to perform and is not usually carried out in
M. Yoshimine routine DSS tests.
Department of Civil Engineering,
Due to anisotropy, the response of sands during loading
Tokyo Metropolitan University, Hachioji,
Tokyo 192-0397, Japan depends on the principal direction. Experimental data show
e-mail: yoshimine-mitsutoshi@c.metro-u.ac.jp that the shear strength and deformation of sands are

123
194 Acta Geotechnica (2009) 4:193–201

dependent on the loading direction. Due to anisotropy, soils conveniently implemented in a spreadsheet environment,
are expected to deform during principal stress rotation even allowing for their ease of use in practice. The model is
when the shear stress level or the mobilized friction angle validated against published experimental DSS and HCT data
is kept constant [1, 8, 13]. on sands.
Another important effect of principal stress rotation is
the non-coincidence or non-coaxiality of the principal
stress and principal plastic strain increment directions. 2 Model for non-coaxial simple shear deformation
Extensive experimental data from Hollow Cylindrical of sand
Testing have shown that the principal stress direction lags
behind the principal plastic strain increment direction when The analytical model for the non-coaxial simple shear
the principal stresses are rotated [1, 7, 8, 13]. The deviation deformation of sand is formulated in two-dimensional
or non-coaxiality between the principal stress and principal plane strain conditions where the stress and strain incre-
plastic strain increment directions is significant for loading ment tensors are defined as
involving pure stress rotations at low mobilized friction    
rx rxy dex dexy
angles. Non-coaxiality has important implications on the rij ¼ ; deij ¼ ð1Þ
rxy ry dexy dey
post-localization response of granular soils as shown by
[8, 25]. Note that dexy = dcxy/2, where dcxy is the engineering shear
Early attempts at modeling simple shear behavior of strain increment. The two-dimensional plane strain
soils were made by [6, 14, 15, 18, 20]. These models did assumption implies that the out-of-plane strains are zero,
not fully account for the effects of principal stress rotation. i.e., dezz ¼ dezx ¼    ¼ 0: All stresses are assumed to be
Recently, models that consider the effects of principal effective, and will be denoted without the symbol prime (0 ).
stress rotation on simple shear response have been pro- In formulating the model, the stress invariants s and t
posed, e.g. [19, 26]. However, most of these models are (which correspond to the mean and shear stresses, respec-
based on complicated theories such as hypo-plasticity, tively, in the plane of deformation) and strain rate invari-
making them difficult to use in routine geotechnical ants dv and dc (which correspond to volumetric and shear
practice. strain rates, respectively) will be used. These invariants are
The objective of this paper is to present a simple consti- defined as follows:
tutive model for the behavior of sands during monotonic  1=2
1 1
simple shear loading that fully accounts for the effects of s ¼ ðrx þ ry Þ; t ¼ ðrx  ry Þ2 þ r2xy ð2Þ
2 4
principal stress rotation. The main feature of the model is the
incorporation of the two important effects of principal stress dv ¼ dex þ dey ; dc ¼ ððdex  dey Þ2 þ 4de2xy Þ1=2 ð3Þ
on stress–strain response: anisotropy and non-coaxiality. In
particular, an anisotropic failure criterion, cross-anisotropic From the Mohr circle for stress (Fig. 1a), it can be seen that
elasticity, and a plastic flow rule and a stress–dilatancy s is equal to the distance to the center of circle from the
relationship that incorporate the effects of principal stress origin, and t is equal to the radius of the circle. Similarly,
rotation are adopted in the model. The model uses simple from the Mohr circle for strain (Fig. 1b), dv/2 is equal to
equations, and the implementation does not require com- the distance to the center of circle from the origin, and dc/2
plicated numerical integration. The model can be easily and is equal to the radius of the circle.

τ

Plane of maximum φ a b
stress ratio
φ dss

(σ , σ )
x xy σ1 direction (dε yy , d γ xy )
dε1 direction
ψ
σ3 2α s σ1 d ε3 2β
σ dv/2

o d ε1
t dγ/2 Zero extension
Pole for (σ y , σ xy )
Pole for ( 0, d γ )
xy
direction
direction direction

Fig. 1 Mohr’s circles: a stress, b incremental strain for simple shear conditions

123
Acta Geotechnica (2009) 4:193–201 195

2.1 Non-coaxial stress–dilatancy relationship Eq. 10 was found to provide a better fit of experimental
simple shear data than Eq. 8, and thus, Eq. 10 will be used
Coaxiality is a fundamental assumption in continuum below in the development of the simple shear model.
mechanics. It implies that the directions of principal plastic In simple shear, the lateral strain is equal to zero,
strain increment and principal stresses coincide. Gutierrez dex = 0. As a result, the volumetric strain increment is
and Ishihara [7] showed that coaxiality does not hold for equal to the vertical strain increment due to dilatancy, i.e.,
loadings involving principal stress rotation, and at the same dvd = deyd, and the dilatancy ratio dvd/dc is equal to
time, conventional plasticity models expressed in terms of deyd
the usual stress and plastic strain increment invariants sin w ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð11Þ
2
deyd þ dc2xy
implicitly assume coaxiality in the plastic flow rule. They
showed that one way to integrate the effects of non-coax-
From the above equation, the simple shear dilatancy ratio
iality in the plastic deformation of granular soils is by
dey/dcxy can be written as
modifying the expressions for energy dissipation and
stress–dilatancy relations. They derived the following deyd sin w
¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ¼ tan w: ð12Þ
stress–dilatancy relationship that is valid for loadings dcxy 1  sin2 w
involving varying degrees of principal stress rotation by
accounting for the degree of non-coaxiality in the calcu- 2.2 Total vertical strain increment
lation of dissipated energy:
In addition to the vertical strain increment from dilatancy
dvd t
¼ sin /c  c ð4Þ deyd, a change in the vertical stress will also induce a
dc s vertical strain deyc due to compression of the soil sample.
where dvd is the volumetric strain increment due to Since the soil sample is constrained from deforming lat-
dilatancy, and /c is the friction angle at phase erally, the compression induced vertical strain increment
transformation, and c is the non-coaxiality factor equal to: deyc is related to the vertical stress increment dry by the
constrained modulus M:
c ¼ cos 2D ð5Þ
dry
The non-coaxiality angle D is equal to the deviation deyc ¼ ð13Þ
M
between the principal stress and the principal strain
increment directions: The total vertical strain increment dey is now equal to

D ¼ ja  bj ð6Þ dry
dey ¼ deyc þ deyd ¼ þ ðtan wÞ dcxy : ð14Þ
M
The angles a and b are shown in Fig. 1 and are defined
2.3 Shear stress–strain relationship
from the Mohr’s circle of stress and strain increment as
2rxy 2dexy A relationship that is widely used in the modeling of the
tan 2a ¼ ; tan 2b ¼ ð7Þ
ry  rx dey  dex non-linear stress–strain response of soils is the hyperbolic
In terms of the dilation angle w and the mobilized stress–strain model. Here, the hyperbolic relation will be
friction angle /, Eq. 4 can also be written as used in terms of the shear stress ratio rxy/ry and the shear
strain cxy:
sin w ¼ sin /c  c sin / ð8Þ
rxy Gmax tanð/dss Þp cxy
dvd t ¼ ð15Þ
sin w ¼ ; sin / ¼ ð9Þ ry tanð/dss Þp þ Gmax cxy
dc s
In addition to the above stress–dilatancy relation derived where (/dss)p is the peak direct simple shear friction angle
from Critical State Soil Mechanics, another widely used corresponding to the peak value of the direct simple shear
stress–dilatancy relation is the one derived by [23]. stress ratio, i.e., tan(/dss)p = (rxy/ry)p and Gmax is the
Gutierrez and Wang [9] recently extended Rowe’s stress– maximum initial shear modulus. The above equation models
dilatancy relationship to account for non-coaxiality. In a shear stress ratio rxy/ry increasing asymptotically towards
terms w and /, the non-coaxial Rowe’s stress–dilatancy the peak shear stress (/dss)p with increasing shear strain cxy.
equation is given as From the Mohr’s circle of stress (Fig. 1), the simple
shear friction angle /dss = tan-1(rxy/ry) can be related to
sin /c  sin / the plane strain friction / = sin-1(r1 - r3)/(r1 ? r3) and
sin w ¼ ð10Þ
cð1  sin / sin /c Þ the principal stress direction a as

123
196 Acta Geotechnica (2009) 4:193–201

sin / sin 2a modeling. It is also worth noting from this figure that
tan /dss ¼ ð16Þ although a full 180° principal stress rotation cannot be
1 þ sin / cos 2a
achieved in simple shear loading, the magnitude of prin-
Note that Eq. 16 is valid for all levels of the mobilized cipal stress is still substantial, approaching as much as 60°.
shear friction angle /, including the peak condition, pro-
vided the corresponding a value is used. 2.5 Non-coaxiality angle in simple shear loading

2.4 Principal stress rotation during simple shear The angle of non-coaxiality D is evaluated using the
loading plastic flow rule developed by [10] for plastic flow in the
(ry - rx)/2 versus rxy stress rotation plane. This flow rule
Oda and Konishi [18], Oda [16], and Ochiai [15] developed is described in Appendix, and gives the following expres-
a simple expression for the amount of principal stress sion for the non-coaxiality angle:
rotation that occurs during simple shear loading. They !
arrived at an expression which reads: 1 1 sin /
D ¼ a  n  sin sinð2a  2nÞ ð18Þ
rxy 2 sin /p
tanð/dss Þ ¼ ¼ j tan a ð17Þ
ry
where / is the mobilized friction angle, /p is the peak
This equation implies a straight-line relationship between friction angle, and n is the principal stress increment
the simple shear stress ratio on a horizontal plane and the direction defined as
tangent of the angle that r1 makes with the vertical axis. 2drxy
The constant j can be shown to be related to the ratio of the tan 2n ¼ : ð19Þ
dry  drx
initial horizontal to the initial vertical stress, which in most
in situ cases is equal to the Ko value. 2.6 Principal stress increment direction
The above equation is compared with the simple shear
experimental data of [4] in Fig. 2. As can be seen, Eq. 17 To obtain an expression for the rotation of the principal
provides a good approximation of the principal stress stress increment direction n, which is required in Eq. 18, it
rotation during simple shear loading of sands. The exper- is only necessary to note that the stress increment vector
imental results show some differences in the degree of should be tangential to the stress path. The stress path is
principal stress rotation between the loose, medium dense given in Eq. 17 in terms of the stress ratio rxy/ry and the
and dense sand samples. However, the effects of density principal stress direction a can be differentiated to obtain
appear to be very small and will be neglected in the this tangent. First, Eq. 17 is re-written in terms of 2a:
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi !
rxy 1 þ tan2 2a  1
¼j ð20Þ
60 ry tan 2a

Substituting Eq. 7 and solving for rx:


50
Princiapal stress direction α (°)

r2xy
rx ¼ þ ð1  jÞry ð21Þ
jry
40
The differential form of the above equation can be written
as
30
2rxy r2xy
drx ¼ drxy  2 dry þ ð1  jÞdry ð22Þ
jry jry
20 Dense
Medium It can be noted that when rxy = 0, drx = (1 - j)dry,
Loose and as a result, it can be concluded that j = (1 - Ko)
10 σxy/σy=κtanα when there is no shear stress applied.
The stress increment drxy can be determined from dcxy
0 and dry using the incremental form of Eq. 15:
0.0 0.2 0.4 0.6 0.8 1.0
Shear stress ratio, σxy /σy Gmax tan/ðdssÞp cxy Gmax ðtan/ðdssÞp Þ2
drxy ¼ dry þ dcxy
tan/ðdssÞp þ Gmax cxy ðtan/ðdssÞp þ Gmax cxy Þ2
Fig. 2 Principal stress rotation during simple shear loading of
Leighton-Buzzard sand (experimental data from [4]) ð23Þ

123
Acta Geotechnica (2009) 4:193–201 197

In case the change in effective vertical stress dry is known


1.1
or prescribed, this value together with drx from Eq. 22 and
drxy from Eq. 23 can be substituted in Eq. 19 to obtain the
principal stress increment direction n. 1.0

sinφp(α)/sinφp(0°)
For drained direct simple shear test, a constant vertical
stress, dry = 0, is usually used, and in this case,
0.9
2rxy drxy
drx ¼ ; dry ¼ 0 ð24Þ
jry Arthur & Menzies (1972)
Oda (1972)
0.8 Tatsuoka (1987)
The principal stress direction n (Eq. 19) can then be Gutierrez (1989)
Model (k=0.87)
written as Model (k=0.94)
Model (k=0.80)
2drxy 0.7
tan 2n ¼  ; dry ¼ 0 ð25Þ
drx 0 15 30 45 60 75 90
Major principal stress direction, α (°)
Combining Eqs. 22 and 24 gives:
ry 1 Fig. 3 Variation of sin /p with major principal stress direction a
tan 2n ¼ j ¼ j ; dry ¼ 0 ð26Þ
rxy tan /dss
and 28 gives the following variation in the peak simple
For undrained loading condition, the change in the shear friction angle (/dss)p with the major principal stress
effective vertical stress dry can be obtained assuming that direction a:
the sample volume is constant during loading, and thus,
2k sin /p ð0 Þ sin 2a
dey = 0. Substituting this condition in Eq. 14 gives the tanð/dss Þp ¼  
vertical stress increment dry in terms of the shear strain ð1 þ kÞ  1  k  2k sin /p ð0 Þ cos 2a
increment during undrained shearing: ð29Þ
dry ¼ Mðtan wÞdcxy : ð27Þ In similar manner, experimental data show that the
elastic deformation of sand is strongly dependent on the
2.7 Anisotropic friction angle and deformation moduli direction of loading. The small strain deformation of sands
has been widely represented by cross-anisotropic elasticity
Extensive experimental data show that the friction angle of where the deformation along the bedding plane is isotropic
sand is strongly dependent on the direction of the applied but different from the deformation perpendicular to the
loading. Studies by [2, 13, 17, 24] showed that the friction bedding plane [3, 5, 11, 12, 28]. For plane strain condition,
angle of sand is highest when the major principal stress is cross-anisotropic elasticity requires four independent
oriented normal to the bedding plane, i.e., when a = 0°, parameters, as shown in the following stress–strain
and decreases as a is rotated from 0° to 90°. Typical relation:
variations of / as function of a from experiments are 8 9 2 38 9
< drx = C11 C12 0 < dex =
shown in Fig. 3.
dry ¼ 4 C12 C22 0 5 dey ð30Þ
The effect of the major principal stress direction on the : ; : dc ;
drxy 0 0 C33 xy
peak friction angle is represented in the model by the fol-
lowing relationship: where C11 = Mh = constrained modulus in the horizontal
2k sin /p ð0 Þ direction, C33 = Mv = constrained modulus in the vertical
sin /p ðaÞ ¼ ð28Þ direction, and C33 = Gvh = shear modulus in the plane of
ð1 þ kÞ  ð1  kÞ cosð2aÞ
anisotropy. Due to the zero lateral strain (dex = 0)
where /p(a) is the direction-dependent friction angle, condition for simple shear loading, only two of these
/p(0°) is the friction angle for a = 0°, and k is the ratio of parameters are required, and these are the vertical
the sine of the friction angles for a = 90° and a = 0°, i.e., constrained modulus Mv = M, and the shear modulus
k = sin /p(90°)/sin /p(0°). The validity of Eq. 28 can be Gvh = Gmax. It can be shown that parameter C12 can be
seen in Fig. 3 where the equation is shown to adequately related to the Ko ratio by the relation:
represent the trend in the anisotropy of the peak friction C12 drx
angle as manifested by the experimental data. The best fit ¼ ¼ Ko ð31Þ
M dry
of Eq. 28 through the experimental data gave a value of
k = 0.84, while the lower and upper bound curves gave To calculate drx, Eq. 31 is only used for consolidation,
values of 0.80 and 0.94, respectively. Combining Eqs. 16 while Eq. 22 is used for shear loading.

123
198 Acta Geotechnica (2009) 4:193–201

3 Comparisons with experimental data Table 1 lists the model parameters used in the simula-
tions of the experimental results. Three of the parameters,
To illustrate its validity, the model described above is used namely the peak friction angle /p, the initial shear modulus
to simulate the DSS test results obtained by [4] on Gmax, and the bulk modulus M are dependent on the initial
Leighton-Buzzard sand, and the HCT results obtained by void ratio eo and the initial vertical stress ryo. For Leigh-
[27] on Kartal sand. Comparisons of model simulations and ton-Buzzard sand, the principal stress rotation parameter j
experimental results are shown in Figs. 4 and 5. The appears to be independent of initial void ratio eo, while for
experimental data for Leighton-Buzzard sand are for Kartal sand j varies with the initial vertical stress ryo. The
drained tests, while the experimental data for Kartal sand dependency of these parameters on eo and ryo are given by
are for undrained tests. Three sets of test results on the expressions in Table 1. The dependency of Gmax on eo
Leighton-Buzzard sand with void ratios at the end of and ryo follows the empirical relation by [21]. The other
consolidation of eo = 0.739, 0.645 and 0.539 are used in parameters, namely the friction angle at phase transfor-
the validation of the model. These void ratios correspond, mation /c and anisotropy parameter k are relatively con-
respectively, to loose, medium dense and dense relative stant and independent of eo and ryo.
densities. All tests were done at constant ry value of As can be seen in Figs. 4 and 5, the simulations of the
100 kPa. For Kartal sand, the test results are for vertical simple shear response are very close to the experimental
consolidation stresses of ryo = 40, 60 and 80 kPa. The data. There is a good agreement between the stress–strain
void ratios at the end of consolidation eo for the three tests response, volumetric response (in drained tests), effective
are very similar ranging from 0.870 to 0.878. stress path (in undrained tests), and the directions of the

0.7 0.8 1.0


loose (eo=0.739) medium dense(eo=0.645) dense (eo=0.529)

Shear stress ratio, σxy/σy


0.6
Shear stress ratio, σxy/σy

Shear stress ratio, σxy/σy

0.8
0.5 0.6

0.4 0.6
0.4
0.3
0.4
0.2
0.2
0.1 Model 0.2
Experiment
0.0
0.0 0.0
0.0 0.2 0.4 0.6 0.0 0.1 0.2 0.3 0.0 0.1 0.2 0.3
Shear strain, γxy Shear strain, γxy Shear strain, γxy
0.020 0.020 0.040
0.015
0.015
Volumetric strain,v

Volumetric strain, v

0.030
Volumetric strain, v

0.010
0.010
0.005 0.020
0.005
0.000
0.010
-0.005 0.000
0.000
-0.010 -0.005
-0.015 -0.010
-0.010
0.0 0.2 0.4 0.6 0.0 0.1 0.2 0.3
0.0 0.1 0.2 0.3
Shear strain, γxy Shear strain, γxy Shear strain, γxy
Direction α, β and ξ (deg.)

Direction α, β and ξ (deg.)

Direction α, β and ξ (deg.)

80 80 80

60 60 60

40 α 40 40
ξ Model
β
20 α 20 20
β Experiment
ξ
0 0 0
0.0 0.2 0.4 0.6 0.0 0.1 0.2 0.3 0.0 0.1 0.2 0.3
Shear strain, γxy Shear strain, γxy Shear strain, γxy

Fig. 4 Comparison of experimental data and model predictions. Top row shear strain versus shear stress, middle row shear strain versus
volumetric strain, and bottom row shear strain versus directions a, b and n (experimental data from [4])

123
Acta Geotechnica (2009) 4:193–201 199

80 80 60
σyo=40 kPa σyo=60 kPa σyo=80 kPa

Shear stress, σxy (kPa)


Shear stress, σ xy (kPa)

Shear stress, σxy (kPa)


60 60
40
40 40

20
20 20
Experiment
Model
0 0 0
0 2 4 6 8 10 0 5 10 15 0 5 10 15
Shear strain, γxy (%) Shear strain, γxy (%) Shear strain, γxy (%)
80 80 60
Shear stress, σxy (kPa)

Shear stress, σ xy (kPa)


Shear stress, σ xy (kPa)
60 60
40
40 40

20
20 20

0 0 0
0 20 40 60 80 0 20 40 60 80 100 0 20 40 60 80 100
Effective vertical stress, σy (kPa) Effective vertical stress, σ y (kPa) Effective vertical stress, σ y (kPa)

30 50 50

σ 1-direction, α (deg)
25
σ1-direction, α (deg)

40 40
σ 1-direction, α (deg.)

20
30 30
15
20 20
10
10 10
5

0 0 0
0 2 4 6 8 10 0 2 4 6 8 10 0 2 4 6 8 10
Shear strain, γxy (%) Shear strain, γxy (%) Shear strain, γxy (%)

Fig. 5 Comparison of experimental data and model predictions. Top row shear strain versus shear stress, middle row effective stress path, and
bottom row shear strain versus principal stress direction a (experimental data from [27])

Table 1 Model parameters used for simulation of simple shear behavior of Leighton-Buzzard sand and Kartal sand
Parameter Leighton-Buzzard sand Kartal sand

Peak friction angle /p for a = 0° (°) 74.6eo - 7.7 82.1(ryo/pa) - 47.7


Friction angle at phase transformation /c (°) 33.5 36.5
Shear modulus Gmax 90F(eo, ryo) 110F(eo, ryo)
Constrained modulus M – 902.7(ryo/pa) - 208.8
Stress rotation parameter j 0.55 2.95 - 2.73(ryo/pa)
Anisotropy parameter k 0.85 0.85
pffiffiffiffiffiffiffiffiffiffiffiffiffi 2
Fðeo ; ryo Þ ¼ pa ryo =pa ð2:97  eo Þ =ð1 þ eo Þ, where eo = initial void ratio, ryo = initial vertical stress, pa = atmospheric pressure

major principal stress a. For the Leighton-Buzzard, the experimental studies by [15, 18, 22] showed that the
predicted and measured principal strain increment direction principal stress rotation is limited (typically to less than
b and the principal stress increment direction n are also in 60°), and that it occurs mainly during small strain defor-
good agreement. mation. Consequently, the effects of principal stress rota-
tion have been routinely neglected in the formulation of the
simple shear response of soils. However, the recent
4 Conclusions emphasis on accurate representation of the pre-failure
deformation of soils has made it essential to account for the
A key feature of the response of soils during simple shear effects of principal stress rotation. To meet this need, the
loading is the rotation of the principal stresses. Early paper presented a constitutive model that specifically

123
200 Acta Geotechnica (2009) 4:193–201

accounts for the effects of principal stress rotation on the increment vector has a length equal to the plastic shear
simple shear behavior of sands. The model incorporated strain increment dc and makes an angle equal to 2b from
two important effects of principal stress on stress–strain the (dey - dex) axis. This plastic strain increment direction
response, namely anisotropy and non-coaxiality. The is evaluated as the normal to the failure surface at the
model used an anisotropic failure criterion, cross-aniso- conjugate point A which is the intersection of the failure
tropic elasticity, and a plastic flow rule and a stress–dilat- surface and the stress increment vector extended from the
ancy relationship that incorporate the effects of non- current stress point. This flow rule is based on the experi-
coaxiality. The model was validated against published mental observations that plastic flow on the (ry - rx)/2
experimental results from direct simple shear and hollow versus rxy stress plane is dependent on the stress increment
cylindrical torsional simple shear tests on sands. direction [10]. This flow rule should be contrasted with
conventional plasticity formulations where the plastic
strain increment direction is evaluated at the current stress
Appendix: Non-coaxiality angle due to principal point independent of the stress increment direction. Details
stress rotation including the experimental verification of the flow rule are
given in [10].
The angle of non-coaxiality angle D required to determine From triangle OAB in Fig. 6, the following angles can
the non-coaxiality parameter c in the stress–dilatancy be obtained:
relationships given in Eqs. 9 and 10 is evaluated using
\OAB ¼ p  ð2n  2aÞ ð33Þ
the flow rule developed by [10] for plastic flow in the
(ry - rx)/2 versus rxy stress rotation plane. This flow rule \BOA ¼ 2D ð34Þ
is shown in Fig. 6. In this figure, the strain increment 2D ¼ p  \OAB  \ABO ð35Þ
components (dey - dex) and 2dexy have been superimposed
on the stress plane. Point A is the current stress point. Using the law of sines:
Neglecting the effect of anisotropy, the failure surface is OA OB
circular and centered at the origin of the (ry - rx)/2 versus ¼ ð36Þ
sin \ABO sin \OAB
rxy axes. The distances OA and OB are equal to the current  
1 OA
shear stress t and the radius of the circular failure surface, \ABO ¼ sin sin \OAB ð37Þ
OB
respectively. These distances can be related to the current
mean stress s, the mobilized friction angle / and the peak Substituting Eqs. 32 and 33 in Eq. 37:
friction angle /p as follows: !
1 sin /
OA ¼ s sin /; OB ¼ s sin /p ð32Þ \ABO ¼ sin sinðp þ 2a  2nÞ ð38Þ
sin /p
On the (ry - rx)/2 versus rxy stress plane, a stress
vector makes an angle equal to 2a from the (ry - rx)/2 Substituting Eqs. 33 and 38 in Eq. 34 gives:
axis (Eq. 7). Similarly, a stress increment vector makes an !
1 sin /
angle equal to 2n from the (ry - rx)/2 axis (Eq. 19). On 2D ¼ 2a  2n  sin sinðp þ 2a  2nÞ ð39Þ
sin /p
the (dey - dex) versus 2dexy strain increment plane, a strain
Simplifying the above equation results in the expression
σ xy 2d ε xy for D given in Eq. 18.
Strain increment
direction

Conjugate stress
B
point
Failure surface References

1. Arthur JRF, Rodriguez JDC, Chua KS, Dunstan T (1980) Prin-


Stress cipal stress rotation: a missing parameter. J Geotech Eng Div
increment ASCE 106(4):419–433
2∆ 2ξ 2. Arthur JRF, Menzies B (1972) Inherent anisotropy in a sand.
A Geotechnique 22:115–128
Current stress
2β 2α point (σ − σ ) / 2
y x
3. Chaudhary SK, Kuwano J, Hayano Y (2004) Measurement of
quasi-elastic stiffness parameters of dense Toyoura sand in hol-
O (dε − dε )
y x low cylindrical apparatus and triaxial apparatus with bender
elements. Geotech Test J 27(1):1–13
Fig. 6 Non-coaxial flow rule in the (ry - rx)/2 versus rxy stress 4. Cole ERL (1967) Soils in the simple shear apparatus. Ph.D.
rotation plane [10] thesis. Cambridge University

123
Acta Geotechnica (2009) 4:193–201 201

5. Fioravante V (2000) Anisotropy of small strain stiffness of Ticino 17. Oda M (1981) Anisotropic strength of cohesionless sands.
and Kenya sands from seismic wave propagation measured in J Geotech Eng ASCE 107(GT9):1219–1231
triaxial testing. Soils Found 40(4):129–142 18. Oda M, Konishi J (1974) Rotation of principal stresses in gran-
6. Frydman S, Talesnick M (1991) Simple shear of isotropic elasto- ular material during simple shear. Soils Found 14(4):39–53
plastic soil. Int J Num Anal Meth Geomech 15(4):251–270 19. Osinov VA, Wu W (2006) Simple shear in sand with an aniso-
7. Gutierrez M, Ishihara K (2000) Non-coaxiality and energy dis- tropic hypoplastic model. Geomech Geoeng 1(1):43–50
sipation in granular materials. Soils Found 40(2):49–59 20. Prevost J, Høeg K (1975) Reanalysis of simple shear soil testing.
8. Gutierrez M, Vardoulakis I (2006) Energy dissipation and post- Can Geotech J 13:418–429
bifurcation behavior of granular soils. Int J Num Anal Meth 21. Richart FE, Hall JR, Woods RD (1970) Vibrations of soils and
Geomech 31(3):435–455 foundations. Int Ser Theor Appl Mech. Prentice-Hall, Englewood
9. Gutierrez M, Wang J (2008) Non-coaxial version of Rowe’s Cliffs, NJ
stress–dilatancy relation. Granural Matter 11:129–137 22. Roscoe KH (1970) The influence of strains in soil mechanics—
10. Gutierrez M, Ishihara K, Towhata I (1993) Flow theory for sand 10th Rankine Lecture. Geotechnique 20(2):129–170
during rotation of principal stress direction. Soils Found 23. Rowe PW (1962) The stress–dilatancy relation for static equi-
31(4):121–132 librium of an assembly of particles in contact. Proc R Soc Lond
11. Jung Y-H, Chung CK, Finno RJ (2004) Development of nonlinear Ser A Math Phys Sci 269(1339):500–527
cross-anisotropic model for the pre-failure deformation of ge- 24. Tatsuoka F (1987) Discussion on the strength and dilatancy of
omaterials. Comp Geotech 31:89–102 sands. Geotechnique 37(2):219–226
12. Kuwano R, Jardine RJ (2002) On the applicability of cross- 25. Vardoulakis I, Georgopoulos IG (2005) The ‘‘stress–dilatancy’’
anisotropic elasticity to granular materials at very small strains. hypothesis revisited: shear-banding related instabilities. Soils
Geotechnique 52:727–749 Found 45(2):61–76
13. Miura K, Miura S, Toki S (1986) Deformation behavior of 26. Yang Y, Yu HS (2006) Numerical simulations of simple shear
anisotropic sand under principal axes rotation. Soils Found with non-coaxial soil models. Int J Num Anal Meth Geomech
26(1):36–52 30(1):1–19
14. Moroto N (1987) On deformation of granular material in simple 27. Yoshimine M, Özay R, Sezen A, Ansal A (1999) Undrained plane
shear. Soils Found 27(1):77–85 strain shear tests on saturated sand using a hollow cylinder tor-
15. Ochiai H (1975) The behaviour of sands in direct shear tests. J sional shear apparatus. Soils Found 39(2):131–136
Jpn Soc Soil Mech Found Eng 15(4):93–100 28. Zeng X, Ni B (1999) Stress-induced anisotropic Gmax of sands
16. Oda M (1975) On the relation s/rn = j tan w in the simple shear and its measurement. J Geotech Geoenviron Eng ASCE 125(9):
test. Soils Found 15(4):35–41 741–749

123

View publication stats

You might also like