Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

MECHANICS

RESEARCH COMMUNICATIONS

Mechanics Research Communications 33 (2006) 515–531


www.elsevier.com/locate/mechrescom

Constitutive modeling of soil-structure interface through


the concept of critical state soil mechanics
a,*
Huabei Liu , Erxiang Song a, Hoe I. Ling b

a
Department of Civil Engineering, Tsinghua University, Beijing 100084, PR China
b
Department of Civil Engineering and Engineering Mechanics, Columbia University, New York, NY 10027, USA

Available online 7 February 2006

Abstract

The behavior of soil-structure interface can be crucial to the overall response of a soil-structure system. The numerical
simulation of soil-structure interaction problem requires proper modeling of the interface. The similarity between the
behavior of soil and interface is first analyzed in the present paper. With this similarity, the concept of critical state soil
mechanics (CSSM), which has been successfully used in the modeling of soil behavior, is used to develop a constitutive
model for soil-structure interface in the framework of generalized plasticity. The model is capable of modeling strain hard-
ening, softening, normal dilatancy and stress-path dependency of interface between sandy soil and structures during shear-
ing. The effects of normal pressure as well as density of sand are captured in the model. The performance of the model is
verified with various experimental results. The unified modeling of the behavior of interfaces with different roughness, dif-
ferent density of soil and different normal pressures using the concept of CSSM is also successfully attempted.
Ó 2006 Elsevier Ltd. All rights reserved.

Keywords: Soil-structure interface; Constitutive model; Critical state soil mechanics; Generalized plasticity

1. Introduction

The serviceability of soil-structure systems, such as shallow foundations, tunnels, and retaining walls,
depends largely on the behavior of the soil-structure interfaces. The mechanical properties of the soil-structure
interface therefore have attracted great attention and many experiments have been conducted in order to gain
insights into the complex phenomena occurred in the interface. Different experimental methods, including
direct shear test, simple shear test, and ring torsional test (e.g., Yoshimi and Kishida, 1981; Desai et al.,
1985; Kishida and Uesugi, 1987; Boulon, 1989; Evgin and Fakharian, 1996), have been employed. These inves-
tigations revealed that the principal factors influencing interface behavior include the roughness of the contact
body, the mineralogical and grading characteristics of soil, the soil density, and the normal pressure, among
which the roughness of the contact body is the most critical one. The up-to-date summary of interface testing
can be found in Hu and Pu (2004). It was also found that a very thin layer of soil adjacent to the contact body

*
Corresponding author. Tel.: +86 10 6278 5681; fax: +86 10 6277 1132.
E-mail address: lhb@tsinghua.edu.cn (H. Liu).

0093-6413/$ - see front matter Ó 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.mechrescom.2006.01.002
516 H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531

together with the roughness of the interface contribute to the shearing behavior of the interface. This thin layer
of soil was termed as the thickness of the interface and was found to be about 5–10 times of the mean diameter
of the soil D50 (Kishida and Uesugi, 1987).
The analogy between the behaviors of sand and rough interface of sandy soil and structures was discussed
by Boulon and Nova (1990). It has been demonstrated that associated with rough interface, the effects of nor-
mal dilatancy, strain hardening, and strain softening for dense interface can be witnessed during shearing. The
density of soil and the normal pressure affect the strength and deformation of interface. Most important of all,
from available experimental results between sandy soil and various rough contact bodies, it can be seen that
the interface reaches an ultimate state at large shearing deformation, at which the stress ratio keeps constant
and the shear deformation continues without dilation or contraction. Additionally, the normal dilation at the
ultimate depends on the normal stress level and the initial void ratio but the ultimate void ratio of the sand
adjacent to the interface depends only on the normal stress level as long as the roughness of the contact body is
the same (Hu and Pu (2004)). These characteristics are similar to those of sand.
The constitutive modeling of soil-structure interface has also been one of the foci in the field of soil-struc-
ture interaction due to the rapid development in the finite element application to the soil-structure analysis.
Apart from constitutive modeling, special interface elements and contact algorithms have also been proposed
for the modeling of soil-structure interface (e.g, Goodman et al., 1968; Katona, 1983; Kaliakin and Li, 1995;
Villard, 1996). However, the modeling of normal dilatancy is difficult using such methods. With the proposal
of thin-layer isoparametric element by Zienkiewicz et al. (1970) and Desai et al. (1984), the use of certain con-
stitutive model to simulate soil-structure interface behavior became possible and various models have been
proposed. Desai et al. (1985) used nonlinear elasticity to model interface behavior, which can simulate the
shearing behavior of interface but was not able to duplicate its normal dilatancy. Aubry et al. (1990), Shahrour
and Rezaie (1997), Ghionna and Mortara (2002), Fakharian and Evgin (2000), Zeghal and Edil (2002),
Gennaro and Frank (2002), Mortara et al. (2002) and others developed constitutive models for soil-structure
interface in the framework of classical elasto-plasticity. Desai and Ma (1992) proposed a model based on the
disturbed-state concept (DSC), and Hu and Pu (2004) used damage mechanics to model soil-structure inter-
face. Most of these models are capable of reproducing the salient behavior of soil-structure interface including
normal dilatancy. However, the analogy between rough interface and soil has not been fully acknowledged.
The objectives of this study are (1) to investigate the similarity between the behavior of sand and interface
and to explore the critical state of interface; (2) to develop a 2D constitutive model in the framework of gen-
eralized plasticity using the concept of critical state soil mechanics; and (3) to unify the modeling soil-structure
interfaces with different roughness within the framework of critical state soil mechanics. The proposed consti-
tutive model was validated against the published experimental results.

2. Basic concepts

Critical state soil mechanics (CSSM) was introduced by the Cambridge Soil Mechanics Group in the 1950’s
and 1960’s (Wood, 1990). Originally, the constitutive models for clays were developed using the concept of
CSSM, the most famous of which were the original and the modified Cam-clay models (Roscoe et al.,
1963; Burland, 1968). With the state parameter as proposed by Been and Jefferies (1985), the CSSM concept
was also successfully employed to describe sand behavior (e.g., Manzari and Dafalias, 1997). The framework
of CSSM states that there exists a critical state of soil at large shearing deformation, and at the critical state
shear deformation continues without dilatancy and change of the stress ratio. The void ratio at this state is the
critical void ratio ec and the stress ratio g = q/p 0 equals to the critical value Mc (where q is the shear stress and
p 0 is the mean effective stress). The critical void ratio ec is affected by the confining pressure such that it
decreases with the increase of confining stress. The behavior of soil at any state thus depends on its distance
between the current state and the critical state, which can be defined by a state parameter for sandy soil.
w ¼ e  ec ð1Þ
Due to the analogy between behaviors of sand and the interface between sandy soils and structures, the
concept of CSSM can be employed to model interface behavior. However, certain modifications need to be
made to the original theory. First of all, the confining stress p 0 in the original theory must be replaced by
H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531 517

the normal stress of the interface rn. Secondly, the stress ratio g = q/p 0 is replaced by g = s/rn, where s is the
shear stress. Since only plane strain problem is considered in the present study, there is only one shear stress in
the tangential direction. Most important of all, the void ratio in the framework of CSSM has to be defined
differently. For an interface, it depends both on the real void ratio of sandy soil and on the roughness of
the contact body. For plane strain problem, neglecting the normal deformation along the interface, the void
ratio increment can be written as:
de ¼ dv=t  ð1 þ eÞ ð2Þ
where t is the thickness of interface, which is about 5–10 times of the mean diameter D50 of sand, and dv is the
deformation increment normal to the interface. Therefore, the critical state void ratio can be determined from
the ultimate normal dilation of interface vult.
Generalized plasticity and its application in geomechanics was firstly introduced by Zienkiewicz and Mroz
(1984) and later extended by Pastor et al. (1990). Several other sand models have been formulated based on
this theory (e.g., Sassa and Sekiguchi, 2001; Ling and Liu, 2003). Within the framework of generalized plas-
ticity, the loading-direction vector {n} and the plastic-flow-direction vector {ng} are defined instead of the
yield and the plastic potential surfaces. The incremental stress tensor can be expressed as:
fdrg ¼ ½Dep fdeg ð3Þ
whereas the elasto-plastic constitutive matrix is given as:

ep e ½De fng gfngT ½De


½D ¼ ½D  T e
ð4Þ
H þ fng ½D ½ng 
in which H is the plastic modulus and [D]e is the elastic constitutive matrix.
Generalized plasticity was used as the framework of the proposed constitutive model. For plane–strain
interface behavior, only the normal and tangential deformations are important, hence resulting in a 2 by 2
elasto-plastic constitutive matrix:
 ep 
ep Dnn Dep ns
½D ¼ ð5Þ
Dep
sn Dep
ss

3. Model description

In the proposed model, the incremental stress and displacement relationship of interface is described
by:
   
drn dv
fdrg ¼ ¼ ½K ð6Þ
ds du
in which drn and ds are the incremental normal and shear stresses, while dv and du are the incremental normal
and tangential displacements, respectively.
With the thickness of interface t, we can define the incremental strains as:
det ¼ du=t; den ¼ dv=t ð7Þ
Hence,
1 ep
fdrg ¼ ½Kfdeg ¼ ½D fdeg ð8Þ
t
in which [D]ep is defined as the elasto-plastic constitutive matrix.
The strains are assumed to be composed of the elastic and plastic portions:
     
den den e den p
fdeg ¼ ¼ þ ð9Þ
det det det
It is assumed that no pure elastic deformation exists in the model.
518 H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531

Nonlinear elasticity is used to determine the elastic deformation. The elastic moduli in the normal and tan-
gential directions are assumed to be uncoupled:
 
e Dn 0
½D ¼ ð10Þ
0 Dt

where Dn and Dt are the normal and tangential moduli, respectively, and the ratio between them R ¼ DDnt is as-
sumed to be constant during shearing.
The tangential modulus is described using an energy-conservative formula (Lade and Nelson, 1987) which
allows the present model to be extended to capture the cyclic behavior of interface:
"   2 #0:5
2
1 þ e rn s
Dt ¼ Dt0 þR
e p0 p0
ð11Þ
Dn Dn0
R¼ ¼
Dt Dt0
in which e is the current void ratio of interface, and p0is reference pressure. p0 can be taken as the atmospheric
pressure. Dn0 and Dt0 are the model parameters. The phrase void ratio of interface is used herein to denote the
density of soil in the thin-layer of soil next to the structure surface. However, as shall be seen in the following,
its value is also affected by the roughness of the interface.
The plastic deformation of interface is determined by theory of generalized plasticity using the concept of
critical state soil mechanics.
First of all, the critical state void ratio is described using the following simple expression:
ec ¼ k lnðrn =p0 Þ þ e0 ð12Þ
where ec is the critical state void ratio, k is the slope of the critical state line in the e  ln p plane, and e0 is the
critical state void ratio at the reference pressure level, which is the atmospheric pressure.
The state parameter w is defined as in Eq. (1) as w = e  ec, where e is the current void ratio.
The loading direction vector is expressed as:
0 1T
B df 1 C
fng ¼ @qffiffiffiffiffiffiffiffiffiffiffiffiffi ; qffiffiffiffiffiffiffiffiffiffiffiffiffiA
2 2
1 þ df 1 þ df ð13Þ
 
af gf
d f ¼ ð1 þ af Þ g
1 þ af

with gf as the stress ratio defining the boundary of possible states. g = s/rn is the current stress ratio and af is a
model parameter. gf depends on the initial state and is assumed to be:
gf ¼ M c þ k hðw0  bÞi ð14Þ
where the Macauley brackets hidefines the operation hxi = x if x P 0 and hxi = 0 if x < 0, while Mc is the crit-
ical state stress ratio s/rn; w0 is the state parameter at initial state; k and b are the model parameters. Eq. (13)
utilizes again the similarity of behavior between soil and interface. Pastor et al. (1990) used the same expres-
sion in the definition of the loading direction vector for soil. It is used here for the interfaces between sandy
soils and structures, but taking into account the influence of the initial state. The parameter b represents the
maximum possible difference in the void ratio of the interface.
Non-associated flow rule is used in the model. According to Ghionna and Mortara (2002), the normal dilat-
ancy of an interface was not a unique straight line. For dense interfaces, there exists a phase transformation
stress ratio Md smaller than the critical state stress ratio Mc, at which the dilatancy is zero. The dense interface
implies that the void ratio of the thin-layer of soil adjacent to the structure surface is small, or in other words,
its relative density is large. The described interface behavior is similar to that of sand. Using this similarity, the
dilatancy of interface is assumed to be a function of stress ratio as follows (Manzari and Dafalias, 1997):
H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531 519

depn
dg ¼ ¼ ag ðM c þ k m w  gÞ ð15Þ
deps
where depn &deps are the incremental plastic strains in the normal and tangential directions, respectively, and ag
and km are the model parameters. Using this expression, associated with dense interface, at a stress ratio smal-
ler than Mc, the dilatancy dg = 0, since the state parameter w < 0 at the state; and at the critical state, w = 0
and the dilatancy remains as zero as g = Mc.
The plastic flow direction is now determined using the following equation:
0 1T
B dg 1 C
fng g ¼ @qffiffiffiffiffiffiffiffiffiffiffiffiffi ; qffiffiffiffiffiffiffiffiffiffiffiffiffiA ð16Þ
2 2
1 þ dg 1 þ dg

The plastic modulus is expressed as:


!
1 g
H ¼ H0 ðrn =p0 Þ 1  ð1 þ gÞ2 ð17Þ
bþw gp

where gp = Mc + kh  wi is the virtual peak stress ratio (Manzari and Dafalias, 1997). H0 is a model
parameter.
From Eq. (17), we can see that the plastic modulus is a function of the state parameter w, with a small value
of w leading to a stiff response of interface. Besides, H = 0 when g = gp and H < 0 when g > gp. gp is a virtual
peak stress ratio because it changes with w and is not the actual peak stress ratio that will be reached by the
current stress ratio g. gp = Mc when w = 0, which leads to H = 0 and from Eq. (15), dg = 0. Thus the critical
state of interface is reproduced. For dense interface, a peak stress ratio gp > Mc will be reached by the current
stress ratio and afterward g P gp, therefore strain softening can also be captured.
The nonlinear normal behavior as described by Desai and Nagaraj (1988) can also reproduced by the plas-
tic modulus in Eq. (17). During normal compression, Eq. (17) reduces to:
1
H ¼ H0 ðrn =p0 Þ ð18Þ
bþw
The plastic modulus increases with the increase in the normal pressure rn and the decrease in the void ratio, as
the state parameter w decreases with the decrease of the void ratio.
The above formulation assumes that the interface is rough. The model can also be extended to capture the
behavior of interfaces with different roughnesses, as will be discussed later.

4. Identification of model parameters

The proposed model requires 11 parameters, all of which have very definite and straightforward physical
meanings. These 11 parameters are: Ds0, Dn0, Mc, k, e0, k, km, af, ag, b, and H0. The identification of the model
parameters can be achieved by conducting two well-controlled constant normal stress shearing tests using the
same contact body and the same sand but with different normal stresses, as well as one normal compression-
test. The thickness of the interface t was implicitly included in the parameters and is equal to 5  10 times of
the mean diameter of the soil D50. In the finite element analysis, it is equal to the thickness of the interface
element (Hu and Pu, 2004). However, for the proposed model, the exact value of the interface thickness
was not important as long as it is properly taken into account in the identification of the model parameters.
The identifications of all the 11 parameters are discussed as follows.

4.1. Elastic parameters: Dt0 and Dn0

The elastic parameters Ds0 and Dn0 can be obtained by observing the initial behavior of the stress–displace-
ment relationship. Using the initial slope of the shear stress–tangential displacement relationship, the initial
tangential stiffness of the interface K it can be obtained. Using the initial slope of the relationship between
520 H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531

the normal stress rn and the normal displacement v, the initial normal stiffness of the interface K in may also be
obtained. The tangential and normal moduli are than obtained using the thickness of the interface t:
Dit ¼ K it =t; Din ¼ K in =t ð19Þ
Since the void ratios of the interface as well as the normal stress rn at the corresponding state are already
known, using Eq. (11), the elastic parameters Dt0 and Dn0 can be readily identified.

4.2. Critical state parameters: Mc, k, and e0

The critical state parameters are identified using the ultimate normal dilation (or contraction) of the inter-
faces as well as the initial void ratio ein and the thickness of the interface t. Suppose two constant normal stress
tests with different normal stresses are conducted with the same initial void ratio of soil. The critical state stress
ratio Mc can be obtained directly using the ultimate stress ratio s/rn of the tests. The initial void ratio of the
interface ein can be taken as similar to that of the soil. The ultimate void ratios of the interface are then
obtained using the following equation:
vult
ec ¼ ein þ ð1 þ ein Þ ð20Þ
t
in which vult is the ultimate normal displacement, ec is the critical state void ratio. The two critical void ratios
corresponding to the two normal stresses are then used to identify k, and e0 in Eq. (12).

4.3. The peak stress ratio parameter: k

The peak stress ratio parameter k is achieved by identifying the peak stress ratio gp and the normal displace-
ment vp when the peak stress ratio is mobilized. Using Eq. (20), the void ratio ep at the peak stress ratio can be
obtained. As the corresponding critical state ec is known, the state parameter wp when the peak stress ratio is
mobilized is obtained using Eq. (1). The parameter k is then identified as:
gp  M c
k¼ ð21Þ
hwp i

4.4. The dilatancy parameters: ag and km

The constant normal stress tests are used to obtain the parameters with the assumptions that the normal
deformation during shearing is purely plastic. The curve of shear stress vs plastic tangential displacement
can be plotted together with the normal deformation vs plastic tangential displacement curve, as illustrated
in Fig. 1. The phase transformation stress ratio Md and the critical state stress ratio Mc may now be directly
read from the figure. The state parameter associated with the phase transformation state wd can also be iden-
tified using the normal deformation. The parameter km can now be obtained using the following expression:
k m ¼ ðM d  M c Þ=wd ð22Þ
The parameter ag is identified using the dilatancy dg at any stress state. For example, the dilatancy d pg at the
peak stress ratio can be used as shown in Fig. 1:
d pg
ag ¼ ð23Þ
M c þ k m w p  gp

4.5. The plastic-modulus parameters: b and H0

The physical meaning of the parameter b is the maximum possible difference in the void ratio of the inter-
face. Usually, b can be taken as:
b ¼ emax  emin ð24Þ
H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531 521

Stress ratio
Mc

Md

Normal displacement
p
d g = dε v /dε s
p p

Plastic tangential displacement

Fig. 1. Identification of the dilatancy parameter a0.

in which emax and emin are the maximum and minimum void ratios of the sand, respectively. Although the void
ratio of the interface depends also on its roughness, according to the authors’ experience, a value of b as shown
in Eq. (24) is a very good initial guess.
The parameter H0 is obtained by matching the normal stress vs normal displacement curve. However, it
could also be obtained directly from Eq. (4) using the normal compression-test. Since it is only 1D, the iden-
tification of H0 using this method is not very intricate. It could also be calibrated by matching the shear stress-
displacement curve, as long as other parameters have already been identified.

4.6. The loading direction parameter: af

The parameter af determines the loading direction. Theoretically, it could be obtained using the contour of
the stress state on the stress space. However, usually a value between 0.4 and 0.5 would be relevant for the
proposed model.
It is commonly accepted that the stress-displacement relationship of a soil-structure interface depends on
the testing methods (e.g., Kishida and Uesugi, 1987). However, from simulation viewpoint, the proposed
model is able to duplicate the experimental results using different testing methods. The differences are mani-
fested in the values of the model parameters. Nevertheless, for the practical application of the model, the
experimental condition of the interface testing must reproduce as much as possible that of the real soil-struc-
ture interaction problem so that the calibrated parameters are reasonable.
The identification of the model parameter assumes that the roughness of the contact bodies used in the tests
is the same as the field problem, and also the density of the soil needs to be properly controlled. As long as
these two conditions are observed, the model parameters corresponding to the specific sand and the contact
body can all be obtained using the aforementioned approaches. However, due to the complexity of the inter-
face behavior as well as the assumptions of the model, as can be seen in the next section, the application of the
model in predicting the results of other interface tests may contain some discrepancies, but within acceptable
limit. The calibration of constitutive model parameters may be based on the regression and optimization
approaches, as discussed by Mattsson et al. (2001) and Yang and Elgamal (2003), although not attempted
in this study.

5. Model evaluation

The experimental results were used to calibrate the proposed model and then the results of simulation were
compared with the experimental results. These test results were obtained from the different testing devices,
including direct shear and simple shear devices. The test conditions covered constant normal stress, constant
normal stiffness and constant volume. The parameters for the various tests are shown in Table 1.
522 H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531

Table 1
Model parameters of the original model
Shahrour and Gennaro and Ghionna and Ghionna and Evgin and Zeghal and
Rezaie (1997) Frank (2002) Mortara (2002)—I: Mortara (2002)—I: Fakharian (1996) Edil (2002)
TiL30 ToD60
Ds0/kPa 300 1000 200 2000 800 800
Dn0/kPa 500 100 200 1500 300 500
Mc 0.65 0.42 0.7 0.6 0.645 0.7
k 0.055 0.25 0.1 0.035 0.07 0.3
e0 0.8 0.75 0.97 0.95 0.995 0.765
k 3.2 4.0 1.5 2.5 0.75 0.5
ag 0.9 0.35 0.9 0.25 0.51 0.9
km 0.5 1.0 0.5 0.5 0.48 0.9
b 0.5 0.4 0.4 0.5 0.3 0.5
H0/kPa 800 80 120 100 350 350
af 0.50 0.50 0.51 0.49 0.52 0.6

5.1. Direct shear tests (Shahrour and Rezaie, 1997)

This set of experiments were constant normal stress tests with different normal stresses and different initial
void ratios. The model parameters were firstly calibrated using the two tests with 100 kPa and 300 kPa normal
stresses but with the same initial void ratio. The thickness of the interface was taken as 4 mm. As no informa-
tion regarding the normal elastic behavior is available, Dn0 was given an arbitrary value of 500 kPa. The same
problem existed for the next simulations of constant normal stress tests and the corresponding normal mod-
ulus was also assigned an arbitrary value.

300 300
n=300 kPa
Model n=200 kPa
Model
250 250
Shear stress (k Pa)

n=100 kPa
Shear stress (kPa)

200 200

150 150

100 100 Dr=90%


Dr=15%

50 50

0 0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
Displacement (mm) Displacement (mm)

0.4 0.2
Vertical displacement (mm)

0.1
Normal displacement (mm)

Model
0.3
0.0

0.2 -0.1
-0.2
0.1
-0.3 Model
=300 kPa
n
=200 kPa
n -0.4 Dr=90%
0.0
=100 kPa
n
Dr=15%
-0.5
-0.1 -0.6
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
(a) Displacement (mm) (b) Displacement (mm)

Fig. 2. Comparison of model prediction and experimental results: interface tests at constant normal stress (Experimental data from
Shahrour and Rezaie, 1997). (a) With different normal stress levels and (b) with different initial void ratios of soil.
H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531 523

The calibrated and the predicted results are shown in Fig. 2 together with the experimental results. It can be
seen that the model is able to reproduce the behavior of soil-structure interface with different normal stresses
and different initial void ratios with one set of parameters. It must be pointed out that the model simulated the
normal dilatancy very satisfactorily, which is very important for soil-structure interaction problem if the inter-
face is restrained in the normal direction.

5.2. Direct shear tests (Gennaro and Frank, 2002)

The constant normal stress tests on the interface between loose Fontainebleau sand (Dr = 46%) and rough
metal plates were simulated, as shown in Fig. 3. The thickness of the interface t was 2 mm. The parameters
were calibrated using the tests with the normal stresses of 50 kPa and 100 kPa, respectively. The test with
25 kPa normal stress was predicted. The model is able to capture the behavior of soil-structure interface at
relatively small normal stress.

5.3. Direct shear tests (Ghionna and Mortara, 2002—I: TiL30)

Two set of experimental results on different interfaces from Ghionna and Mortara (2002) were simulated,
namely TiL30 and ToD60. The model parameters for interface TiL30 were calibrated using the results of
the constant normal stiffness tests. The two tests with k = 0 and k = 100 kPa were used to identify the
parameters. The normal elastic modulus was obtained by matching the normal stress vs the tangential dis-
placement curve of the constant stiffness tests. Thickness of the interface t was taken as 2 mm. Fig. 4 shows
the results of calibration as well as those of the prediction. The prediction agreed with the experimental
results very well.

100
n =100 kPa
n =50 kPa
80 n =25 kPa
Shear stress (kPa)

60
Model
40

20

0
0 1 2 3 4 5 6
Tangential displacement (mm)

0.4
n =100 kPa
0.3
Vertical displacement (mm)

=50 kPa Model


n
0.2 n =25 kPa

0.1
0.0
-0.1
-0.2
-0.3
-0.4
0 1 2 3 4 5 6 7
Tangential displacement (mm)

Fig. 3. Comparison of model prediction and experimental results: interface tests at constant normal stress (Experimental data from
Gennaro and Frank, 2002).
524 H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531

200
TiL30, n0
=50 kPa

160

Shear stress (kPa)


120 Model k=0 kPa/mm
k=100 kPa/mm
k=1000 kPa/mm
80

40

0
0 2 4 6 8 10
Tangential displacement (mm)

TiL30, =50 kPa


250 n0
Normal stress (kPa)

200
Model
k=0 kPa/mm
150 k=100 kPa/mm
k=1000 kPa/mm

100

50

0 2 4 6 8 10
Tangential displacement (mm)

Fig. 4. Comparison of model prediction and experimental results: interface tests at constant normal stiffness—TiL30 (Experimental data
from Ghionna and Mortara, 2002).

5.4. Direct shear tests (Ghionna and Mortara, 2002—I: ToD60)

The model parameters for interface ToD60 were calibrated using the constant stiffness tests with
k = 1000 kPa and rn0 = 100 kPa and rn0 = 300 kPa (t = 1.5 mm). The model parameters were then used
to predict the behavior of the interface with k = 1000 kPa and rn0 = 200 kPa as well as the constant normal
stress tests. As the normal deformations of the tests were not available, only the comparison of the stress–
tangential displacement relationship for the constant normal stress test and the stress path curves of
the constant stiffness tests are presented in Fig. 5. It can be seen that the predicted results are satisfac-
tory.

5.5. Simple shear tests (Evgin and Fakharian, 1996)

The stress-path tests by Evgin and Fakharian (1996) on sand/steel interface were predicted. The thickness
of the interface t was 3 mm. The model parameters were first identified using the constant normal stress tests,
the results of which are shown Fig. 6(a). The model parameters were then used to predict the results of the
constant stiffness results, as shown in Fig. 6(b) and (c). The model duplicated satisfactorily the experimental
results.

5.6. Direct shear tests (Zeghal and Edil, 2002)

The model’s capability in simulating constant volume test was verified using the direct shear tests of Zeghal
and Edil (2002). The comparison between the experimental and simulated results is shown in Fig. 7
H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531 525

400
ToD60, k=1000 kPa/mm
350
n0 =100 kPa

300 n0 =200 kPa

Shear stress (kPa)


n0 =300 kPa
250

200

150 Model

100

50

0
0 100 200 300 400 500 600
Normal stress (kPa)

350
ToD60, k=0 kPa/mm
n0 =100 kPa
300 Model
n0 =200 kPa

n0 =300 kPa
Shear stress (kPa)

250

200

150

100

50

0
0 1 2 3 4 5 6 7
Tangential displacement (mm)

Fig. 5. Comparison of model prediction and experimental results: interface tests at constant normal stiffness—ToD60 (Experimental data
from Ghionna and Mortara, 2002).

(t = 2 mm). The model is capable of modeling both the constant normal stress test and the constant volume
test.

6. Unification of interface behaviors with various roughnesses

Up to now, the proposed model is able to reproduce the salient behavior of rough interfaces under different
normal stresses and with various initial void ratios. The model can also be extended to describe the different
behaviors of interfaces with various roughnesses.

6.1. Interface behaviors with various roughnesses

According to the experimental results by Uesugi and Kishida (1986), Kishida and Uesugi (1987), and Hu
and Pu (2004), the following conclusions can be made regarding the behaviors of interfaces with different
roughnesses.

(1) The roughness of interface can be quantified using the normalized roughness Rn (Kishida and Uesugi,
1987):
Rn ¼ Rmax ðL ¼ D50 Þ=D50 ð25Þ
in which Rmax(L = D50) is the Rmax value of a contact body surface with gage length L = D50, and D50 is
the mean diameter of soil in the interface.
526 H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531

500 0.6
σn =100 kPa
Model

Normal displacement (mm)


0.5
σn =300 kPa
400
σn =500 kPa 0.4
Model
Shear stress

300 0.3

0.2
200 0.1 σn =500 kPa
0.0 σn =300 kPa
100 σn =100 kPa
-0.1

0 -0.2
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Tangential displacement (mm) Tangential displacement (mm)
(a)
500 0.4

Normal displacement (mm)


0.3
400
Shear stress (kPa)

0.2
300 σn0 =100 kPa
0.1 Model
σn0 =200 kPa
200 σn0 =300 kPa
0.0
σn0 =100 kPa

100
Model σn0 =200 kPa
-0.1
σn0 =300 kPa

0 -0.2
0 1 2 3 4 5 6 0 1 2 3 4 5 6

Tangential displacement (mm) Tangential displacement (mm)

600 400
σn0 =100 kPa
350 σn0 =200 kPa
500
σn0 =300 kPa
Normal stress (kPa)

300
Shear stress (kPa)

400
250

300 200

Model σn0 =100 kPa 150 Model


200
σn0 =200 kPa
100
σn0 =300 kPa
100
50

0 0
0 1 2 3 4 5 6 0 100 200 300 400 500 600
Tangential displacement (mm) Normal stress (kPa)
(b)

Fig. 6. Comparison of model prediction and experimental results: stress path testing (experimental data from Evgin and Fakharian, 1996).
(a) Constant normal stress tests, (b) Constant normal Stiffness test (k = 800 kPa/mm), (c) with different normal stiffnesses rn0.

(2) Soil-structure interfaces can be differentiated as rough and smooth interfaces using a critical normalized
roughness Rcr. When Rn P Rcr, the interface can be regarded as rough. The yielding of smooth interfaces
is almost elastic-perfectly plastic and the normal dilation can be neglected.
H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531 527

300 400
k=0 kPa/mm
k=200 kPa/mm σn0=100 kPa 350 σn0=100 kPa
250 k=400 kPa/mm
k=600 kPa/mm 300

Normal stress (kPa)


Shear stress (kPa)

k=800 kPa/mm
200
250

150 200

150
100 k=200 kPa/mm
k=400 kPa/mm
100
k=600 kPa/mm
50 k=800 kPa/mm
50

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Tangential displacement (mm) Tangential displacement (mm)

k=0 kPa/mm σn0=100 kPa


0.6 k=200 kPa/mm
Model
Normal displacement (mm)

k=400 kPa/mm
k=600 kPa/mm
0.4 k=800 kPa/mm

0.2

0.0

-0.2
0 1 2 3 4 5 6
Tangential displacement (mm)
(c)

Fig. 6. (continued)

(3) The relationship between the stress ratios at yield and the normalized roughness Rn for smooth interface
is approximately linear, while the critical stress ratios of rough interfaces are almost the same; the peak
stress ratio of rough interface increases with the increase of the normalized roughness Rn.
(4) The normal dilation at the critical state for rough interfaces increases with the increase of the normalized
roughness Rn.

6.2. Extension of proposed model to capture the interface behaviors with various roughnesses

Using the aforementioned experimental facts, it can be seen that the framework of CSSM is suitable in
describing the behaviors of interfaces with different roughnesses. The capability of proposed model can be
extended through the following modifications.
First of all, the critical state stress ratio Mc is set as constant when Rn P Rcr. When Rn < Rcr, it is expressed
as a function of Rn:

M c ¼ M ac þ M bc Rn ð26Þ

Secondly, the critical state void ratio e0 at the atmospheric pressure associated rough interface is assumed to
be:
528 H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531

60 120
Experimental result
50 100 Model prediction

Shear stress (kPa)


Shear stress (kPa)

40 80

30 60
Experimental result

20 Model prediction 40

10 20

0 0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 0 1 2 3 4 5

Tangential displacement (mm) Tangential displacement (mm)

0.20 150
Experimental result
Experimental result
Normal displacement (mm)

0.15 Model prediction


Model prediction 125
Normal stress (kPa)
0.10
100
0.05
75
0.00

50
-0.05

-0.10 25
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 0 1 2 3 4 5
(a) Tangential displacement (mm) (b) Tangential displacement (mm)

Fig. 7. Comparison of model prediction and experimental results (experimental data from Zeghal and Edil, 2002). (a) Constant normal
stress test and (b) constant volume test.

e0 ¼ ea0 þ eb0 Rn ð27Þ


If the interface is smooth, e0 is assumed to be a constant and the critical normalized roughness Rcr is substi-
tuted into Eq. (27) to obtain the corresponding value.
The dilatancy coefficient ag must also depend on the roughness of the interface. For smooth interface
(Rn < Rcr), ag = 0; associated with rough interface, again a linear function is used to describe the dependency:
ag ¼ aag þ abg Rn ð28Þ
In order to model the dependency of the peak stress ratio on the roughness, the parameter k is also assumed
to be a linear function of Rn when Rn P Rcr:
k ¼ k a þ k b Rn ð29Þ
If Rn < Rcr, k = 0.
With these modifications, the model is able to capture the different interface behaviors with various rough-
nesses. For verification, the experimental results by Hu and Pu (2004) were simulated using the modified
model. The parameters are shown in Table 2, the number of which is five more than that of the original
one. The thickness of the interface was taken as 5 mm, which is 5 times of the Ds0 of the sand used in the tests.
The comparison between the experimental results and the predictions is shown in Fig. 8. The model duplicated
H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531 529

Table 2
Model parameters of the extended model
Rcr Ds0/kPa Dn0/kPa M ac M bc k ea0 eb0
0.1 9000 3000 0.1778 4.222 0.06 0.99 0.155
aag abg ka kb km b H0/kPa af
0.5 0.3 0.55 1.8 0.6 0.5 900 0.6

250
Rn=0.01
Model Rn=0.05
200
Rn=0.10
Shear stress (kPa)

Rn=0.20
150 Rn=0.50

100

50

0
0 2 4 6 8 10
Tangential displacement (mm)

1.0 Rn=0.01
Rn=0.05
Normal displacement (mm)

0.8 Rn=0.10
Rn=0.20
0.6 Rn=0.50

Model
0.4

0.2

0.0

0 2 4 6 8 10
Tangential displacement (mm)

Fig. 8. Comparison of model prediction and experimental results: with various roughnesses (experimental data from Hu and Pu, 2004).

the experimental results to a very acceptable extent. The calibration procedure of interface roughness can be
found in Kishida and Uesugi (1987), and the Rn’s used in the simulation of the test results were obtained
directly from Hu and Pu (2004).
Simple linear functions were used in the aforementioned modification. More complicated relationships may
also be used for some of the model parameters, but it will add to the complexity of the model and it can be
seen from the simulation of experimental results that it is not necessary.
At this stage, the modified model unifies the behaviors of the interfaces between sandy soil and structure. As
long as the sands are the same and the contact bodies are made of the same material, the complicated behaviors
of the soil-structure interfaces can be captured using the proposed model with one single set of model para-
meters. A total of 16 parameters may seem too many. However, the calibration of them is very straightforward
and easy to accomplish. It must also be pointed out that for practical application such as pile foundation,
530 H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531

usually the original model with 11 parameters is enough since the roughnesses of the contact bodies are approxi-
mately the same.

7. Conclusions

In the present study, the behavior of the interface between sandy soil and rigid structures was investigated
and the analogy between the behaviors of sands and rough interfaces was analyzed. Based on the analogy, a
new constitutive model in the framework of generalized plasticity was proposed using the concept of critical
state soil mechanics (CSSM). The basic features of sand-structure interface could be described by the proposed
model. These include: hardening/softening mechanical response, phase transformation, critical state, and
dependency of dilatancy and friction on normal stress level and the initial void ratio. The model parameters
have direct physical meaning and can be identified rather easily following a series of procedures. The capabil-
ity of the model in simulating the experimental results of various tests, including stress-path tests, indicated
that the concept of critical state soil mechanics could be used to model the interface behavior. The proposed
model was also extended successfully to unify the interface behaviors with different roughnesses. The extended
model can simulate the complicated behavior of sand-structure interfaces with different roughnesses, different
initial void ratios, and different normal stress levels with one single set of model parameters. It is also believed
that improvements of the present version of constitutive model to include cyclic loading could be achieved
without major difficulties.

Acknowledgements

The authors would like to express their appreciations for the support offered by the National Science Foun-
dation of China (Grant No. 50378050) and the Science Foundation of Beijing (Grant No. 8011002).

References

Aubry, D., Modaressi, A., Modaressi, H., 1990. A constitutive model for cyclic behavior of interfaces with variable dilatancy. Comput.
Geotech. 9, 47–58.
Been, K., Jefferies, M.G., 1985. A state parameter for sands. Geotechnique 35 (2), 99–112.
Boulon, M., 1989. Basic features of soil structure interface behavior. Comput. Geotech. 7, 115–131.
Boulon, M., Nova, R., 1990. Modeling soil-structure interface behavior: a comparison between elastoplastic and rate type laws. Comput.
Geotech. 9, 21–46.
Burland, J.B., 1968. Correspondence on ‘The yielding and dilation of clay’. Geotechnique 15, 211–214.
Desai, C., Drumm, C., Zaman, M., 1985. Cyclic testing and modeling of interfaces. J. Geotech. Eng. ASCE 111 (GT6), 793–815.
Desai, C.S., Ma, Y., 1992. Modeling of joints and interfaces using the disturbed-state concept. Int. J. Numer. Anal. Meth. Geomech. 16,
623–653.
Desai, C.S., Nagaraj, B.K., 1998. Modeling for cyclic normal and shear behavior of interfaces. J. Eng. Mech. ASCE, 114(7), pp. 1199–
1217.
Desai, C.S., Zaman, M.M., Lightner, J.G., Siriwardane, H.J., 1984. Thin-layer element for interface and joints. Int. J. Numer. Anal. Meth.
Geomech. 8, 19–43.
Evgin, E., Fakharian, K., 1996. Effect of stress paths on the behavior of sand-steel interfaces. Can. Geotech. J. 33, 853–865.
Fakharian, K., Evgin, E., 2000. Elasto-plastic modeling of stress-path-dependent behavior of interfaces. Int. J. Numer. Anal. Meth.
Geomech. 24, 183–199.
Gennaro, V., Frank, R., 2002. Elasto-plastic analysis of the interface behavior between granular media and structure. Comput. Geotech.
29, 547–572.
Ghionna, V., Mortara, G., 2002. An elastoplastic model for sand-structure interface behavior. Geotechnique 52 (1), 41–50.
Goodman, R.E., Taylor, R.L., Brekke, T.L., 1968. Model for mechanics of jointed rock. J. Soil Mech. Found. Div. 94 (SM3), 637–659.
Hu, L.M., Pu, J.L., 2004. Testing and modeling of soil-structure interface. J. Geotech. Geoenviron. Eng. ASCE 130 (8), 851–860.
Kaliakin, V., Li, J., 1995. Insight into deficiencies associated with commonly used zero-thickness interface elements. Comput. Geotech. 17,
225–252.
Katona, M.G., 1983. A simple contact friction interface element with applications to buried culverts. Int. J. Numer. Anal. Meth.
Geomech. 7, 371–384.
Kishida, H., Uesugi, M., 1987. Tests of interfaces between sand and steel in the simple shear apparatus. Geotechnique 37 (1), 45–52.
Lade, P., Nelson, R., 1987. Modeling of the elastic behavior of granular materials. Int. J. Numer. Anal. Meth. Geomech. 11, 521–542.
Ling, H., Liu, H., 2003. Pressure dependency and densification behavior of sand through a generalized plasticity model. J. Eng. Mech.
ASCE 129 (8), 851–860.
H. Liu et al. / Mechanics Research Communications 33 (2006) 515–531 531

Manzari, M., Dafalias, Y., 1997. A critical state two-surface plasticity model for sands. Geotechnique 47 (2), 255–272.
Mattsson, H., Klisinski, M., Axelsson, K., 2001. Optimization routine for identification of model parameters in soil plasticity. Int. J.
Numer. Anal. Meth. Geomech. 25 (5), 435–472.
Mortara, G., Boulon, M., Ghionna, V., 2002. A 2-D constitutive model for cyclic interface behavior. Int. J. Numer. Anal. Meth.
Geomech. 26, 1071–1096.
Pastor, M., Zienkiewicz, O.C., Chan, A.H.C., 1990. Generalized plasticity and the modeling of soil behavior. Int. J. Numer. Anal. Meth.
Geomech. 14 (3), 151–190.
Roscoe, K.H., Schofield, A.N., Thuraijajah, A., 1963. Yielding of clays in states wetter than critical. Geotechnique 13, 211–240.
Sassa, S., Sekiguchi, H., 2001. Analysis of waved-induced liquefaction of sand beds. Geotechnique 51 (2), 115–126.
Shahrour, I., Rezaie, F., 1997. An elastoplastic constitutive relation for the soil-structure interface under cyclic loading. Comput. Geotech.
21 (1), 21–39.
Uesugi, M., Kishida, H., 1986. Frictional resistance at yield between dry sand and mild steel. Soils Found. 26 (4), 139–149.
Villard, P., 1996. Modelling of interface problems by the finite element method with considerable displacements. Comput. Geotech. 19 (1),
23–45.
Wood, D., 1990. Soil Behavior and Critical State Soil Mechanics. Cambridge University Press, Cambridge.
Yang, Z.H., Elgamal, A., 2003. Application of unconstrained optimization and sensitivity analysis to calibration of a soil constitutive
model. Int. J. Numer. Anal. Meth. Geomech. 27 (15), 1277–1297.
Yoshimi, Y., Kishida, T., 1981. A ring torsion apparatus for evaluating friction between soil and metal surface. Geotech. Test. J. ASTM 4
(4), 145–152.
Zeghal, M., Edil, T., 2002. Soil structure interaction analysis: modeling the interface. Can. Geotech. J. 39, 620–628.
Zienkiewicz, O.C., Best, B., Dulllage, C., Stagg, K.G., 1970. Analysis of nonlinear problems with particular reference to jointed rock
systems. In: Proc. of 2nd Int. Congress on Society of Rock Mechanics 3, 501–509.
Zienkiewicz, O.C., Mroz, Z., 1984. Generalized plasticity formulation and application to geomechanics. In: Desai, C.S., Gallangher, R.H.
(Eds.), Mechanics of Engineering Materials. John Wiley and Sons, New York, pp. 655–680.

You might also like