Você está na página 1de 20

Review

Review of fundamental principles in modelling unsaturated soil behaviour


Daichao Sheng

Centre for Geotechnical and Materials Modelling, School of Engineering, The University of Newcastle, NSW 2308, Australia
a r t i c l e i n f o
Article history:
Received 5 February 2011
Received in revised form 18 April 2011
Accepted 9 May 2011
Keywords:
Unsaturated soils
Constitutive modelling
Volume change
Shear strength
Yield stress
Water retention
Hydro-mechanical coupling
a b s t r a c t
An unsaturated soil is a state of the soil. All soils can be partially saturated with water. Therefore, consti-
tutive models for soils should ideally represent the soil behaviour over entire ranges of possible pore
pressure and stress values and allow arbitrary stress and hydraulic paths within these ranges. The last
two decades or so have seen signicant advances in modelling unsaturated soil behaviour. This paper
presents a review of constitutive models for unsaturated soils. In particular, it focuses on the fundamental
principles that govern the volume change, shear strength, yield stress, water retention and hydro-
mechanical coupling. Alternative forms of these principles are critically examined in terms of their pre-
dictive capacity for experimental data, the consistency between these principles and the continuity
between saturated and unsaturated states.
2011 Elsevier Ltd. All rights reserved.
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 757
2. Net stress and suction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 758
3. Volume change behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 759
3.1. Separate stress and suction approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 759
3.2. Combined stresssuction approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 760
3.3. SFG approach. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 762
3.4. Stress-path dependent elastic behaviour . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 763
4. Yield stress versus suction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 763
4.1. Relationship between yield stress and volume change equation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 763
4.2. Reconstituted soils versus compacted soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 766
5. Shear strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 768
6. Water retention behaviour and hydro-mechanical coupling. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 769
7. Finite element implementation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 772
8. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 773
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 773
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 773
1. Introduction
Soils that are partially saturated with water are often referred to
as unsaturated soils. Some soils exhibit distinctive volume,
strength and hydraulic properties when become unsaturated. For
these soils, a change in the degree of saturation can cause signi-
cant changes in volume, shear strength and hydraulic properties.
Nevertheless, the distinctive volume, strength and hydraulic
behaviour for unsaturated states should be treated as material
nonlinearity and modelled consistently and coherently. In other
words, a constitutive model for a soil should represent the soil
0266-352X/$ - see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compgeo.2011.05.002

Tel.: +61 2 49215746; fax: +61 2 49216991.


E-mail address: daichao.sheng@newcastle.edu.au
Computers and Geotechnics 38 (2011) 757776
Contents lists available at ScienceDirect
Computers and Geotechnics
j our nal homepage: www. el sevi er . com/ l ocat e/ compgeo
behaviour over entire ranges of possible pore pressure and stress
values and should allow arbitrary stress and hydraulic paths with-
in these ranges. After all, any soil can be partially saturated with
water and an unsaturated soil is only a state of the soil, not a
new soil, as pointed out by Gens et al. [36].
Soil mechanics principles are more established for soils at satu-
rated states. Generalisation of these principles to unsaturated soils
requires careful consideration of these fundamental issues: (1) vol-
ume change behaviour associated with suction or saturation
changes, (2) shear strength behaviour associated with suction or
saturation changes, and (3) hydraulic behaviour associated with
suction or saturation changes. Soils can experience signicant vol-
ume changes upon changes of the degree of saturation or suction.
Some soils expand upon wetting, some collapse and some do both
depending on the stress level. The large volume changes associated
with saturation change can lead to severe damages to foundations
and structures. Shear strength of soils can also change dramatically
as the degree of saturation changes, and a related engineering
problem is slope failures caused by rainfall. Unsaturated soils also
have distinctive hydraulic behaviour which has profound implica-
tions in designing cover and containment systems for various
industrial and municipal wastes. These fundamental issues are in-
deed the main concerns of unsaturated soil mechanics and its engi-
neering applications [26,45].
Constitutive modelling of unsaturated soils generally involves
the generalisation of constitutive models for saturated states to
unsaturated states, by incorporating the fundamental issues men-
tioned above. Research in this direction was pioneered by Alonso
et al. [2] and it has since attracted extensive interest. A large num-
ber of constitutive models can now be found in the literature.
There are several state-of-the-art reviews over the last 15 years
or so, e.g. Gens [32], Wheeler and Karube [148], Kohgo [58], Wheel-
er [146], Gens et al. [37], Sheng and Fredlund [107], Sheng et al.
[110], Gens [33], Cui and Sun [19] and Gens [34]. These papers
may serve as good references for studying the topic. They usually
provide (1) a thorough discussion of stress state or constitutive
variables used to establish various models, (2) an in-depth analysis
of specic constitutive models and their advantages and disadvan-
tages, and (3) latest developments in the area of unsaturated soil
modelling. The papers by Gens [33,34] also provide interesting dis-
cussions on the physical signicance of different suction compo-
nents and their roles in constitutive modelling.
Instead of a comprehensive review of existing constitutive mod-
els for unsaturated soils, this paper devotes its main attention to a
number of specic issues: (1) volume change behaviour, (2) varia-
tion of yield stress and shear strength with suction, (3) water
retention behaviour and hydro-mechanical coupling. These issues
represent the most fundamental components of constitutive mod-
els for unsaturated soils. Alternative methods for tackling these
fundamental issues are scrutinised, particularly against the princi-
ple that partial saturation is a state of soil and against their predic-
tive capacity for experimental data. The issue of stress state
variables, a seemingly unavoidable topic in unsaturated soil
mechanics, is not specically discussed in this paper. In other
words, constitutive models are not judged based on the stress state
variables they use, rather on their qualitative predictions of ob-
served soil behaviour. In addition, a number of advanced topics
are not discussed in this paper:
Thermodynamics of unsaturated soils. This topic often leads to
interesting discussion of thermodynamically consistent stress-
state variables and constitutive models. Interested readers may re-
fer to, e.g. Houlsby [44], Hutter et al. [47], Gray and Schreer [41],
Sheng et al. [114], Li [63], Samat et al. [100], Coussy et al. [16] and
Zhao et al. [157].
Micromechanical modelling of unsaturated soils. This approach
can often lead to insights into soil behaviour and sometimes also
validation of macroscopic (continuum) constitutive equations.
Some good examples of this approach include the work by Gili
and Alonso [39], Jiang et al. [49], Katti et al. [52] and Scholts
et al. [104]. A closely related approach is the micromacro dou-
ble-structure models by Gens and Alonso [35], Alonso et al. [4],
Snchez et al. [101] and Cardoso and Alonso [13].
Thermo-hydro-chemo-mechanical modelling of unsaturated
soils. In this approach, additional environmental variables such
as temperature and chemical concentrations are introduced to
study soil behaviour. Some recent work in this area refers to Loret
et al. [67], Guimars et al. [43], Cleall et al. [15], Kimoto et al. [56],
Gens et al. [36,37] and Gens [34].
Miscellaneous topics such as non-isothermal behaviour, anisot-
ropy, rate-dependent behaviour, degradation and damage, lique-
faction of unsaturated soils. These topics are usually related to
specic soils or specic problems. Some representative work on
these topics are Cui and Delage [17], Stropeit et al. [121], Romero
and Jommi [97] and DOnza et al. [24] for anisotropy; Modaressi
and Modaressi [77] and Cui et al. [18] for non-isothermal behav-
iour; Cardoso and Alonso [13] for degradation modelling; Arson
and Gatmiri [5] and Yang et al. [151] for damage modelling; Olde-
cop and Alonso [85] and Pereira and de Gennaro [91] for rate-
dependent behaviour; Unno et al. [138] and Bian and Shahrour
[7] for liquefaction.
The paper is organised as follows. It rst presents a brief discus-
sion of two basic concepts used in constitutive modelling, i.e. the
net stress and the matric suction. The fundamental issues on vol-
ume change, yield stress, shear strength, water retention and hy-
dro-mechanical coupling are then discussed. The paper nally
outlines the challenges and possible solution strategies for imple-
menting unsaturated soil models into the nite element method.
2. Net stress and suction
Net stress is commonly used in interpreting unsaturated soil
behaviour and in constitutive modelling. It is dened as
r
ij
r
ij
d
ij
u
a
1
where r
ij
is the net stress tensor, r
ij
the total stress tensor, d
ij
the
Kronecker delta, and u
a
the pore air pressure. The net stress is often
used to analyse laboratory data, particularly those based on the
axis-translation technique where the air pressure is not zero. It is
sometimes perceived to recover the effective stress when soils be-
come saturated. Such a perception should however be avoided. Un-
der natural ground conditions where the air pressure is
atmospheric, the net stress is equivalent to the total stress. Indeed,
we never use the net stress concept to describe the behaviour of dry
sand. In other words, the atmospheric pore air pressure should be
considered zero. Net stress is different from total stress only when
the air pressure is not atmospheric.
The concept of net stress can be useful in interpreting experi-
mental data based on axis-translation technique, if the technique
is indeed valid for applying suction (see discussions on its validity
in e.g. [80,6]). In this case, the air pressure is used as a reference
value for stress measures and the net stress is the total stress in ex-
cess of the pore air pressure.
The soil suction in the literature of unsaturated soil modelling
usually refers to the matric suction and is usually expressed as:
s u
a
u
w
2
where s is called the matric suction in soil physics terminology and
also called the matrix suction [6], and u
w
the pore water pressure.
The matric suction is used interchangeably with the matric poten-
tial in soilwater potential. The latter is a measurement of energy
and consists of two parts: capillary and adsorptive potentials. When
758 D. Sheng / Computers and Geotechnics 38 (2011) 757776
pore water exists as capillary water at relatively high degrees of sat-
uration, the capillary potential is dominant in the matric potential
and the denition by (2) is then considered to be valid. However,
when pore water exists as adsorbed water lms at low degrees of
saturation, the adsorptive potential (w
a
) becomes dominant in the
matric potential. Consequently, questions have been raised regard-
ing whether Eq. (2) is still valid for the matric potential [6]. In the
case when water exists as adsorbed lms to solid particles, the true
water pressure is not well dened. It is not unique at one material
point and is dependent on the proximity to the solid particle surface
[70]. However an apparent water pressure can be introduced:
u
w
= u
a
w
a
, i.e. the apparent water pressure represents the nega-
tive adsorptive potential measured in excess of air pressure. Such
an apparent water pressure is then unique at one material point.
With such a denition of u
w
, the matric potential can be expressed
by Eq. (2) over the full range of saturation. Nevertheless, the matric
suction should be differentiated from capillary phenomena. Its two
components may not easily be separable in a soil with double-
porosity. As pointed out in Gens [34], it is more appropriate to think
of matric suction as a variable that expresses quantitatively the de-
gree of attachment of water to solid particles that results from the
general solid/water/interface interaction.
In constitutive modelling, the matric suction is often treated as
an additional variable in a stress space for establishing constitutive
laws. This approach was pioneered in the Barcelona Basic Model
(BBM) by Alonso et al. [2] and is then followed in most existing
models, with a few exceptions where the suction is treated as an
internal variable or hardening parameter (e.g. [9,68]). Again, these
approaches are not differentiated in this paper. Since the matric
suction can vary independently of stress, it is treated as an inde-
pendent axis in the stress space in this paper. In addition, the mat-
ric suction is considered to coincide with the negative pore water
pressure for fully saturated states and thus the suction axis extends
from negative innity to positive innity in the stress space.
3. Volume change behaviour
The volume change behaviour is one of the most fundamental
properties of soils. For unsaturated soils, the large volume changes
associated with suction change can cause severe damages to foun-
dations and structures. The volume change equation also under-
pins the yield stresssuction and shear strengthsuction
relationships [110]. It is indeed the only absolutely necessary com-
ponent that is needed to extend a saturated soil model to unsatu-
rated states. The model that denes the volume change caused by
stress and suction changes should again be applicable to the entire
range of possible pore pressure or suction values. The discussions
below are limited to isotropic stress states. The volume change
associated with changes of deviator or shear stress has to be con-
sidered in a three-dimensional constitutive framework, which de-
pends on the specic model used for saturated soils, and is outside
the scope of this paper.
For saturated soils, a common starting point is the linear rela-
tionship between the specic volume (v) and the logarithmic effec-
tive mean stress (ln p
0
) for normally consolidated soils:
v N k lnp
0
N k lnp u
w
3
where p is the mean stress, k is the slope of the v ln p
0
line, and N
is the intercept on the v axis when ln p
0
= 0. Eq. (3) is only valid for
positive increments of the effective stress. For unloading and
reloading, the volume change depends on the specic plasticity
framework adopted in the constitutive model. For example, hypo-
plasticity and bounding surface plasticity adopts different volume
change mechanisms than classical elastoplasticity. However, for
normally consolidated soils subject to positive stress increments,
Eq. (3) is usually used independently of the theoretical framework.
It should be noted that Eq. (3) represents a straight line in the
v ln p space only if the pore water pressure is zero. If the pore
water pressure were kept at a negative value (suction), equation
(3) would predict a smooth curve in the v ln p space, as shown
by Fig. 1. The air entry suction for the soil in Fig. 1 is assumed to
be larger than 100 kPa, so that the soil remains saturated. Indeed,
these compression lines look very much like those for overconsol-
idated soils. However, the curvature of the normal compression
lines is purely due to the nature of the logarithmic function and
the translation from the effective stress space (v ln p
0
) to the total
stress space (v ln p), not due to overconsolidation.
Equation (3) can also be written in an incremental form as
follows:
dv k
dp
p u
w
k
du
w

p u
w
4
It is clear that a negative increment in pore water pressure has ex-
actly the same effect on the volume change of a saturated soil as an
equal positive increment in mean stress.
In the literature, Eq. (3) is extended to unsaturated states in one
of the three approaches:
Approach A: Separate stress and suction approach, or the net
stress and suction approach.
Approach B: Combined stresssuction approach, or the effective
stress approach.
Approach C: SFG approach, which is a middle ground between
Approach A and B.
These approaches are discussed separately below in terms of
their advantages and disadvantages.
3.1. Separate stress and suction approach
In Approach A, the volume change due to stress change is sep-
arated from that due to suction change. A typical example of the
volume change equations in this approach is:
v N k
vp
ln p k
vs
ln
s u
at
u
at
_ _
5
where p is the net mean stress, N is the specic volume when
ln p 0 and s = 0, k
vp
is the slope of an assumed v ln p line or
the compressibility due to stress change, k
vs
the slope of an as-
sumed v ln s line or the shrinkability due to suction change, and
1.5
2.0
2.5
3.0
1 10 100 1000
v
p, kPa
u
w
=0
u
w
=-10 kPa
u
w
=-100 kPa
Air entry value larger than 100 kPa
Fig. 1. Normal compression lines for saturated clay under constant pore water
pressures (k = 0.2, N = 3).
D. Sheng / Computers and Geotechnics 38 (2011) 757776 759
u
at
is the atmospheric pressure and is added to avoid the singularity
when s = 0. Again, Eq. (5) is only used for increasing mean stress or
increasing suction. Indeed, k
vs
is usually replaced by the elastic
compression index j
vs
in most applications, unless the suction in-
creases to a historically high value (above the suction-increase yield
surface, see section below on yield stress).
Eq. (5) has been used in many models such as Alonso et al. [2],
Wheeler and Sivakumar [150], Cui and Delage [17], Chiu & Ng [14],
Georgiadis et al. [38] and Thu et al. [134]. The main advantage of
Eq. (5) is that the compressibility due to stress and suction changes
are dealt with separately. This does not only provide extra exibil-
ity for modelling soil behaviour, but is also supported by experi-
mental data. Toll [136] and Toll and Ong [137] showed that the
two compressibilities k
vp
and k
vs
can be totally different (Fig. 2).
It is usually true that the suction shrinkability (k
vs
) decreases with
decreasing degree of saturation. On the other hand, the stress com-
pressibility (k
vp
) can increase with increasing suction, particularly
for compacted soils where highly compressible macropores
(inter-aggregate pores) are present [96,30].
However, there are also a few shortcomings about Eq. (5). First,
Eq. (5) does not recover Eq. (3) when the soil becomes saturated. It
represents a linear v ln p relationship for constant suctions, un-
less k
vp
is assumed to be a function of stress (as in [38]. This is
not consistent with the saturated soil model (Fig. 1). As a conse-
quence, the volume change becomes undened at the transition
suction between saturated and unsaturated states. Second, the
volume change caused by suction changes is independent of stress,
which is at variance with experimental observation shown in
Fig. 3. In addition, the atmospheric pressure (u
at
) in (5) makes
the suction change insignicant when the suction is less than the
atmospheric pressure (s < u
at
). The rst point, i.e. the discontinuity
at the transition between saturated and unsaturated states, was
also one of the reasons that some researchers turned to the effec-
tive stress approach (e.g. [112]). A simple numerical example will
illustrate this problem. Let a soil be compressed at the transition
suction (s
sa
) from mean stress 1 kPa to 100 kPa. Let the air pressure
remain atmospheric. In the saturated zone, the volume changes
according to Eq. (3):
Dvj
s

sa
k
vp
ln
100 s
sa
1 s
sa
In the unsaturated zone, the volume changes according to Eq. (5):
Dvj
s

sa
k
vp
ln100
These two volume changes can be quite different, depending on the
value of the transition suction. The transition suction is either the
air-entry or the air-expulsion value, depending on the hydraulic
path.
3.2. Combined stresssuction approach
In Approach B, the matric suction and the net mean stress are
combined into one single variable, i.e. an effective stress, to dene
their effects on soil volume. A general form of the effective stress is
p
0
p f s 6
where f is either a function of suction or a function of suction and
degree of saturation. Such a denition of effective stress is very gen-
eral and covers most existing denitions in the literature (but per-
haps not the recent one by [144]. It is noted that function f(s)
usually involves material state variables such as the degree of satu-
ration and the air entry value. As a consequence, the denition of
the effective stress and hence the stress space can change with
material states, a feature which is not shared by its counterpart
for saturated soils.
With such an effective stress, Eq. (3) is usually assumed to be
still valid for unsaturated states:
v N k lnp
0
N ks ln p f s 7
where N is the specic volume when ln p
0
= 0. If the effective stress
is indeed effective in controlling soil volume, v should remain con-
stant under constant p
0
. As such, parameters N and k should be inde-
pendent of suction. However, this is seldom the case in most
effective stress models. In the literature, k is usually assumed to
be function of s, while N is treated either as a constant or a variable.
We rst discuss the case where N is a constant and then show that a
varying N with suction cannot be recommended.
Eq. (7) is widely used in the literature and in fact most effective
stress models adopt it as the volume change equation (e.g.
[9,12,50,57,59,68,82,83,92,103,112,114,123,125]). Eq. (7) generally
recovers Eq. (3) when the soil becomes saturated. This is one of the
greatest advantages of using the effective stress.
However, there are also some disadvantages with Eq. (7). The
obvious issue is the difculty in addressing the different compress-
bilities due to stress and suction changes, as shown in Fig. 2, since
there is now only one compressibility in the volume change equa-
tion. The second issue is related to a constraint on the compress-
ibility k. Let a saturated slurry soil be dried from zero suction to
an arbitrary suction under constant mean stress of 1 kPa, i.e. the
stress path AB in Fig. 4. As will be shown in the section on yield
(S
r
=1)

vp

vs
S
r
(%)

Fig. 2. Variation of compressibilities with degree of saturation (after [136]).


E1 NC 25 kPa
E7 NC 200 kPa
E11 NC 400 kPa
E2 NC 50 kPa
E5 OC 200 kPa
E6 OC 800 kPa
1 10 100 1000
Suction (kPa)
V
o
i
d

r
a
t
i
o

0.9
0.8
0.7
0.6
0.5
0.4
Fig. 3. Variation of shrinkability with stress (after [23]).
760 D. Sheng / Computers and Geotechnics 38 (2011) 757776
stress (Fig. 9), this drying path is elastoplastic, not purely elastic.
The volume of the soil then changes according to:
v
B
N ks ln1 f s 8
Now compress the soil under constant suction, i.e. stress path BC in
Fig. 4. The compression line will be curved in the v ln p space, due
to the f(s) term. If the suction at point B is above the air entry value,
this compression line is expected to intersect with the initial com-
pression line for saturated states. Let the intersection be point C (at
net mean stress of p
r
). The volume at C is then:
v
C
v
B
ks ln
p
r
f s
1 f s
_ _
9
Along path AC, the volume changes according to:
v
C
N k0 ln p
r
f 0 N k0 ln p
r
10
Replacing (8) into (9) and then equating (9) with (10) lead to:
ks
k0

ln p
r

ln p
r
f s
< 1 11
We usually anticipate the effective stress to increase with increas-
ing suction, at least at low suction values. Therefore, we have:
k(s) < k (0), meaning that the slope of the compression line de-
creases with increasing suction. Such a constraint on k is however
not supported by experimental data. In Alonso et al. [2], the slope
of the compression lines decreases with increasing suction. In the
data by Toll [136], Sivakumar and Wheeler [117], Toll and Ong
[137] and Gallipoli et al. [30] for compacted soils, the slope of the
compression lines increases with increasing suction. In addition,
data on wetting-induced collapse (e.g. [126]) do not support an ever
increasing collapse volume with increasing mean stress.
In some constitutive models that use the combined stresssuc-
tion approach, parameter N is assumed to vary with suction. If N
decreases with increasing suction, the same constraint on k, i.e.
Eq. (11) would apply. To avoid this constraint, N has to increase
with increasing suction. However, the variation of N with suction
has implications on the yield surface evolution. An increasing N
with suction basically implies that drying a slurry soil under con-
stant effective mean stress will cause the yield surface to expand,
which is not consistent with the stress path. This inconsistency will
be discussed in association with Fig. 9 below.
Gallipoli et al. [30] proposed the following volume change
equation:
v N k lnp
0
1 a1 expbn 12
where N and k are the two parameters of the normal compression
line for saturated states, a and b tting parameters, and n a positive
variable representing the bonding effects of suction. The bonding
variable (n) is a function of both s and S
r
. Eq. (12) is hence equivalent
to Eq. (7) with N and k being functions of s and S
r
. Gallipoli et al.
[29,30] showed that Eq. (12) is able to predict the volume change
at both normal compression and critical states for a variety of com-
pacted soils. One challenge in using Eq. (12) is that the yield stress is
likely to be functions of both suction and degree of saturation. Be-
cause the sS
r
relationship is usually not unique due to hydraulic
hysteresis, the resulting loading-collapse surface may not be well
dened. However, Gallipoli (personal communication) suggests that
it is sufcient to dene the bonding variable (n) in terms of S
r
only
(e.g. n = 1 S
r
), which would then resolve the non-uniqueness
problem and lead to a unique loading-collapse surface in the S
r
p
0
space.
To avoid the constraint dened by Eq. (11), a possible augmen-
tation to Eq. (7) is to assume that the compressibility k is a function
of degree of saturation, i.e. k(S
r
), while keeping N constant. This is
similar to the approach by Gallipoli et al. [30] with n being a func-
tion of S
r
only. Because S
r
changes with soil volume even if the suc-
tion is kept constant, the slope of the compression line will then
change and will most likely increase with increasing mean stress.
Therefore, it is possible to have k(S
r
) decreasing with decreasing
S
r
, but the slope of the compression line for constant suction in-
creases with increasing stress (Fig. 5b). Al-Badran and Schanz [1]
used an approach similar to Fig. 5b, but formulated their volume
change equation in the net stresssuction space. Kikumoto et al.
[55] and Zhang and Ikariya [152] assumed a saturation-dependent
compressibility (k(S
r
)), but formulated their models in the effective
stresssuction space. The same issue of non-uniqueness of the
yield surface in the stresssuction space will arise, due to hydraulic
hysteresis. To avoid this non-uniqueness problem, the yield surface
and hence the shear strength can be dened in the S
r
p
0
space,
thus eliminating suction as the additional variable of the stress
space. Some work in this direction has recently been reported by
Zhou et al. [159].
Clearly further research is required if the combined stress
suction approach is adopted for the volume change, particularly in
terms of consistent explanation of the suction-caused and stress-
caused volume changes for reconstituted soils. A worthwhile
endeavour in this direction is perhaps to explore the possibilities
1 p
r
N
ln p
A
B
C
s=0
s>0, ds=0
v
Drying path constant p
N
ln p
A
B
s=0
s>0, ds=0
v
Drying path under constant p ?
C
(a) (b)
Fig. 4. Normal compression lines according to the effective stress approach constraint on k(s) with a constant N: (a) net mean stress space, (b) effective mean stress space.
D. Sheng / Computers and Geotechnics 38 (2011) 757776 761
of using k(S
r
) in Eq. (7) and the S
r
p
0
space to establish all consti-
tutive equations.
3.3. SFG approach
Sheng et al. [108] proposed a third way to model the volume
change for unsaturated soils under isotropic stress states. The
new model, referred to as the SFG model, represents a middle
ground between Approach A and Approach B and is expressed in
an incremental form as follows:
dv k
vp
d p
p f s
k
vs
ds
p f s
13
Eq. (13) is in the same form as Eq. (4). Similar to Approach A, Eq.
(13) is dened in terms of net stress and suction and separates
the compressibilities due to the two variables, i.e. k
vp
and k
vs
. Sim-
ilar to Approach B, it combines the suction with the net mean stress
in the denominator, i.e. the term p f s, and recovers Eq. (4) for
saturated states. The term p f s represents the interaction be-
tween stress and suction and makes the normal compression lines
for non-zero suction curved in the v ln p space. However, there
is no constraint on parameter k
vp
. As a rst approximation, k
vp
can be assumed to be independent of suction, as indicated by the
data of Jennings and Burland [48] for air-dry soils. More realistically
it should depend on suction. For example, the data of Sivakumar
and Wheeler [117] shows that k
vp
increases with increasing suction
for compacted soils. Parameter k
vs
must equal k
vp
when the soil is
fully saturated, because of Eq. (4). It generally decreases with
increasing suction and approaches zero. Sheng et al. [108] sug-
gested the following simple function for k
vs
k
vs

k
vp
; s 6 s
sa
k
vp
ssa1
s1
; s > s
sa
_
14
where s
sa
is the transition suction and was also called the saturation
suction in Sheng et al. [108]. It is the unique transition suction be-
tween saturated and unsaturated states in the SFG model.
We note that the number 1 in Eq. (14) is used to avoid the sin-
gularity when s
sa
= 0 and is not truly needed if s
sa
is not absolutely
zero. A better expression would be:
k
vs

k
vp
; s 6 s
sa
k
vp
ssa
s
; s > s
sa
_
15
The difference between Eqs. (14) and (15) is minimal, but Eq. (15) is
preferred. Eq. (15) can be applied as long as the transition suction is
not absolutely zero. Kurucuk et al. [61] used Eq. (13), but a different
function for k
vs
than (15). Again, both k
vp
and k
vs
can vary with
stress path and take different values on a loading and unloading
path respectively.
The function f(s) in Eq. (13) can also take different forms. Sheng
et al. [108] initially used the following function:
f s s 16
This is perhaps the simplest form possible for f(s) and yet guaran-
tees the continuity between saturated and unsaturated states. Even
with this simplest form, Zhou and Sheng [158] showed that the SFG
model is able to predict a good set of experimental data on volume
change and shear strength, both for soils reconstituted from slurry
and for compacted soils. Due to the f(s) term in Eq. (13), the normal
compression lines will be curved in the v ln p space (Fig. 6). Fig. 6a
shows the compression curves for a reconstituted soil under various
suctions. The compression curves for s = 400, 650 1000 kPa are all
normal compression lines that do not involve any unloading or
reloading. The SFG predictions for these curves were obtained with
one single k
vp
value. Fig. 6a also shows the reconstituted soil be-
comes stiffer as the suction increases. Fig. 6b shows the compres-
sion curves for a compacted soil. The difference in the estimated
yield stresses by Eqs. (5) and (13) is clearly shown in the gure
for s = 300 kPa. The yield stress is indicated by the meeting points
of the unloadingreloading line and the normal compression lines.
The parameter k
vp
was allowed to change with suction in Fig. 6b.
One shortcoming of Eq. (16) is that the soil compressibility ap-
proaches zero as suction increases to innite. There is also a theo-
retical discontinuity between unsaturated states and completely
dry state (S
r
= 0). To avoid these problems, an alternative form of
f(s) could be used, for example:
f s S
r
s 17
This equation will not only guarantee the continuity between satu-
rated and unsaturated states, but also the continuity between
unsaturated and the completely dry state (S
r
= 0). More interest-
ingly, both Eqs. (16) and (17) will lead to the same shear
strengthsuction relationship. The degree of saturation in (17) can
also be replaced by the effective degree of saturation (S
e
r
), as sug-
gested by Pereira and Alonso [90] when discussing Bishops v
parameter. However, the performance of Eq. (17) is yet to be
1
ln p
A
NCL(s>0, ds=0)
v
s=0
( )
r
S
N
NCL under constant
r
S
1
N(s)
ln p
A
NCL(s=0)
NCL(s>0, ds=0)
v
Drying path under constant p ?
N(0)
( )
r
, s S
Drying path under constant p ?
(a) (b)
Fig. 5. Other options of normal compression lines according to the effective stress approach: (a) N increases with increasing suction. (b) k is a function of degree of saturation
(S
r
).
762 D. Sheng / Computers and Geotechnics 38 (2011) 757776
validated against experimental data, and the loading-collapse yield
surface may become non-unique due to hydraulic hysteresis.
The SFG approach seems to be able to overcome some disadvan-
tages of Approach A and Approach B. The main disadvantage of the
SFG approach is that it exists only in an incremental form and its
integration depends on stress path [154,109]. The stress path
dependency requires special treatment in the stress integration
of the constitutive model [111].
3.4. Stress-path dependent elastic behaviour
An interesting issue related to the volume change is that all the
approaches discussed above are stress-path dependent. Zhang and
Lytton [154] and Sheng et al. [109] showed that Approaches A and
C lead to stress-path dependent elastic behaviour. Approach B can
also result in stress-path dependent elastic behaviour, because a
closed loop of net stress and suction changes do not necessarily
lead to a closed loop of effective stress changes (Fig. 7). Such
behaviour is caused by the material state dependency of the effec-
tive stress. For example, the change of the effective mean stress
along path AB in Fig. 7 is usually different from that along path
CD, because the degree of saturation (or the air entry value) has
changed from A to C, even though the net mean stress is kept con-
stant during both paths. Such an unclosed loop means that the
model is stress-path dependent, even if the stress changes occur
inside the elastic zone. A model that exhibits stress-path depen-
dent elastic behaviour is at variance with classical elastoplasticity
theory and thermodynamics, and hence should generally be
avoided. However, it is still a challenge to develop a model that
exhibits a stress-path independent elastic behaviour.
4. Yield stress versus suction
4.1. Relationship between yield stress and volume change equation
In the literature of unsaturated soil mechanics, the yield stress
of an unsaturated soil is usually assumed to be a function of suc-
tion. The concept of yield stress in classical elastoplasticity theory
refers to the stress level that causes plastic deformation. In other
plasticity frameworks such as bounding surface plasticity
[21,99,78], hypoplasticity [60,72] and generalised plasticity
[69,86,101], plastic deformation occurs along all loading paths
including reloading. In these cases, a loading function or a bound-
ing surface is used to differentiate unloading from loading. In the
discussion below, the yield surface concept is based on the classical
elastoplasticity theory, but can be extended to the loading or
bounding surface concepts.
Under isotropic stress states, the yield stress is also called the
preconsolidation stress. For unsaturated soils, the yield net mean
stress, denoted here by p
c
, is conventionally determined from iso-
tropic compression curves obtained at constant suctions. These
compression curves are usually plotted in the space of void ratio
versus logarithmic net mean stress. The initial portion of the curve
is typically atter than the ending portion of the curve, if the suc-
tion is larger than zero. Such a curve is then approximated by two
straight lines, one representing the elastic unloadingreloading
line and the other the elastoplastic normal compression line. The
intersection point of the two straight lines is used to dene the
preconsolidation (yield) stress (Fig. 8a). The yield stress so deter-
mined is generally found to increase with increasing suction, irre-
spective of samples prepared from slurry states or from compacted
soils, leading to the so-called loading-collapse yield surface that is
widely used in constitutive models for unsaturated soils (Fig. 8b).
The procedure outlined above for determining the yield stress is
based on the assumption that the e ln p relationship for normally
consolidated soils at s > 0 is linear and may not be consistent with
the denition of yield stress. To demonstrate this inconsistency, we
should rst realise that the isotropic compression curves shown in
Fig. 8 are typical of unsaturated soils reconstituted from slurry (e.g.
v
(b) Compacted Kaolin (a) Air-dry silty clay
1 10 100 1000 10000
1.54
1.50
1.46
1.42
1.38
s = 0 kPa
s = 400 kPa
s = 650 kPa
s =1000 kPa
Measured Predicted by eq. (13)
p (kPa)
Reloading line
NCL
NCL
NCLs
p (kPa)
NCL fitted by
eq. (13)
s = 0
s = 300 kPa
Reloading
2.25
2.15
2.05
1.95
10 100 1000
NCL fitted by eq. (5)
v
Reloading
Fig. 6. Predicted isotropic compression curves for (a) reconstituted silty clay (data by [20]), (b) compacted kaolin (data by [117]).
A
B
C
D
p
s s
A
B
C
D
p
D
Fig. 7. A closed loop in the net stress space and its corresponding open loop in the
effective stress space.
D. Sheng / Computers and Geotechnics 38 (2011) 757776 763
[48]) as well as of compacted soils. Because it is easier to under-
stand the preconsolidation stress for a slurry soil that for a com-
pacted soil, we use a slurry soil as an example here. Let us
assume that the slurry soil has not been consolidated (with a zero
preconsolidation stress). The initial yield stress for the soil is then
zero (Point A in Fig. 9a). We also note that the effective stress for
saturated states is constant along the 135 line in the p s space.
Drying the slurry soil to suction B under zero mean stress is similar
to consolidating the soil to stress E under zero suction (Fig. 9a). In-
deed, if the air entry value of the soil is larger than suction at B, the
length AB is exactly the same as AE, due to the effective stress prin-
ciple for saturated soils. However if the air entry value is lower
than suction at B, the length AB should generally be larger than
AE, because a suction increment is generally less effective than
s
p
Current yield
surfaces
45
o
A
B
C
D
F E G
135
o
45
o
A
B
C D
F
G
E
p
s
Current yield
surfaces
Z
e
r
o

s
h
e
a
r

s
t
r
e
n
g
t
h

l
i
n
e

(a) Net stress suction space
H
Z
e
r
o

s
h
e
a
r

s
t
r
e
n
g
t
h

l
i
n
e

c0
p
c0
p
v
A
B
C
D
ln p
NCL (s>0, ds=0)
NCL (s=0)
(c) Corresponding volume change in net
stress space
(d) Corresponding volume change in effective
stress space
N
ln p
A, H
NCL(s=0)
NCL(s>0, ds=0) with (S
r
)
v
Drying path under constant
net mean stress
C
D
B
(b) Effective stress suction space
Fig. 9. Evolution of the elastic zone during drying and compression of a slurry soil (A ?B ?C: constant net mean stress).
(b) Variation of yield stress with suction (a) Isotropic compression curves
ln p
c1
p
c2
p
c3
p
0 s =
s
1
1 2 3
0 s s s < < <
c1 c2 c3
p p p < <
s
2
s
3
e
s
p
s
1
s
2
s
3
c1
p
c2
p
c3
p
LC yield surface
Fig. 8. Isotropic compression curves under constant suction (s) and derived yield stresses.
764 D. Sheng / Computers and Geotechnics 38 (2011) 757776
an equal stress increment in terms of consolidating the unsatu-
rated soil. Once the soil becomes unsaturated, the yield stress does
not necessarily change with suction along the 135 line (denoted
by the dashed line). However, the new yield surface always go
through the current stress point, as shown in Fig. 9a. The elastic
zone expands as the suction increases from A to B. The stress points
(e.g. B and C) are on the current yield surfaces. Let now the soil be
isotropically compressed under the constant suction at Point C
(stress path CD in Fig. 9a). The isotropic compression path (CD) is
again outside the initial elastic zone and the soil at point C is nor-
mally consolidated. Therefore, the isotropic compression path (CD)
as well as the drying path from B to C is elastoplastic and does not
involve a purely elastic portion as Fig. 8 indicates, suggesting that
the method for determining the yield stress in Fig. 8 be conceptu-
ally inconsistent. Indeed, the suction-induced apparent consolida-
tion effect should refer to the increase of the preconsolidation
stress at zero suction ( p
c
0 moves from E to F as suction increases
from B to C in Fig. 9a), not the preconsolidation stress at the cur-
rent suction. We also note that as the slurry soil is dried under con-
stant stress, the soil becomes stiffer. The slope of the compression
line at point C is much smaller than that at point A in Fig. 9c.
The same analysis can be done in the effective stresssuction
space (Fig. 9b). In the p
0
s space, the zero shear strength line or
the apparent tensile strength surface is commonly assumed to
be: p
0
= 0, i.e. the vertical line that goes through the origin of the
space. This line must also be the initial yield surface for a soil that
has not been consolidated, i.e. p
0
c
0 0. For a saturated soil that
has a zero yield stress, i.e. a slurry, the size of the elastic zone re-
mains zero as long as the effective mean stress remains zero, irre-
spective of the pore water pressure values. When the soil becomes
unsaturated, the yield surface will continue along the zero shear
strength line, if the effective mean stress remains zero. To keep
the effective mean stress zero, a tensile net mean stress has to be
applied to balance out the suction increase. As a consequence,
the size of the elastic zone remains zero, which also reects the
effective stress principle. Furthermore, if the yield surface does
not collapse to the zero shear strength line when p
0
c
0 0, there
would be plastic deformation for loading along the yield surface,
and the shear strength as well as the yield stress of the soil would
become non-uniquely dened.
In the effective stresssuction space (Fig. 9b), the stress path for
suction increase under constant net mean stress is initially inclined
to horizontal by 45 for saturated states. Therefore, the stress path
will cross the current yield surface and the drying path AB for a
slurry soil is elastoplastic, not purely elastic. Once the soil becomes
unsaturated, the stress path will drift away from the 45 line and
the yield surface will also drift away from the vertical line (the
dashed lines in Fig. 9b). Nevertheless, the current stress points A,
B, C, and D stay on the current yield surfaces and the stress path
ABCD causes elastoplastic volume change, in consistency with
Fig. 9a.
Fig. 9b can also be used to illustrate that parameter N in the vol-
ume change equation, i.e. Eq. (7), cannot be a function of suction.
This is because increasing suction under constant effective mean
stress, i.e. stress path AH in Fig. 9b, does not lead to the expansion
of the elastic zone. To obtain stress path AH in Fig. 9b, a tensile net
mean stress has to be applied as suction increases, so that the
effective mean stress remains zero. Such a stress path corresponds
to the so-called neutral loading and does not lead to yield surface
expansion. The expansion of the elastic zone is a prerequisite for
shifting the normal compression lines towards higher specic vol-
ume in v ln p
0
space (Fig. 5a). Therefore, parameter N in Eq. (7)
cannot be a function of suction, if the zero shear strength line is as-
sumed to be p
0
= 0. The corresponding volume change along path
ABCD can still be explained consistently if the soil compression in-
dex (k) is assumed to be a function of degree of saturation, as
shown in Fig. 9d. Points A and H are then identical in the v ln p
0
space.
In the literature, the yield stress variation with suction is a
rather confusing part. Many early models adopt three yield sur-
faces in the net stresssuction space, namely the loading-collapse
yield surface, the suction-increase yield surface and the apparent
tensile strength surface. Fig. 10 shows some examples of these
models. The loading-collapse yield surface is used to model the
volume collapse when an unsaturated soil is rst loaded under
constant suction and then wetted under constant stress. The
suction-increase yield surface is used to capture the plastic volume
change when an unsaturated soil is dried to a historically high suc-
tion. The apparent tensile strength surface denes the zero shear
strength or the apparent tensile strength due to suction increase.
These yield surfaces are usually dened separately. For example,
setting the preconsolidation stress ( p
c0
) to zero in the loading-
collapse yield function does not recover the apparent tensile
strength function. The suction-increase yield surface is usually
horizontal or gently sloped (Fig. 10) and is not related to the
loading-collapse surface or the apparent tensile strength surface.
Sheng et al. [108] showed that the loading-collapse surface, the
apparent tensile strength surface and the suction-increase yield
surface are related to each other. In the SFG model, the yield
stresssuction relationship, the apparent tensile strengthsuction
relationship and the shear strengthsuction relationship are all de-
rived from the volume change equation, i.e. Eq. (13). In this model,
the yield stress for a slurry soil that has never been consolidated or
dried varies with suction in a unique function. This function also
denes the apparent tensile strength surface or the zero shear
strength surface in the stresssuction space (the curve through
point A in Fig. 11). The curve approaches the 45 line as the suction
becomes zero or negative (positive pore water pressure). Drying
this slurry soil under zero stress (stress path ABC) causes the
expansion of the yield surface to point C in Fig. 10. Therefore, the
suction-increase yielding is already included in the yield function
and there is no need to dene a separate function. If the unsatu-
rated soil at point C is then compressed under constant suction
(stress path CD in Fig. 11), the yield surface will evolve to the load-
ing-collapse surface that passes through point D in Fig. 11. The
yield surface in the stress space represents the contours of the
hardening parameter, which is usually the plastic volumetric
strain. The stress path CD will change the initial shape of the yield
surface, because the plastic volumetric strain along CD depends on
the suction level. The loading-collapse yield function recovers the
apparent tensile strength function when the preconsolidation
p
c0
p
Suction-increase
yield surfaces
Loading collapse
yield surfaces
Apparent
tensile strength
surface
Thu et al. (2007a)
Alonso et al. (1990)
Delage and Graham (1995)
s
Fig. 10. Loading-collapse, suction-increase and apparent tensile strength surfaces
in various models.
D. Sheng / Computers and Geotechnics 38 (2011) 757776 765
stress at zero suction ( p
c0
) is set to zero. All these yield surfaces are
continuous and smooth in the stresssuction space. Models based
on effective stresses also have these features of the SFG model, as
shown in Fig. 9b.
As pointed out by Wheeler and Karube [148] and shown by
Sheng et al. [108] and Zhang and Lytton [156], the apparent tensile
strength function, the suction-increase yield function and the load-
ing-collapse yield function are all related to the volumetric model
that denes the elastic and elastoplastic volume changes caused by
stress and suction changes. If Approach A, i.e. Eq. (5), is adopted to
describe the volume change, the loading-collapse yield surface will
take the following form:
p
c

p
c0
s; s 6 s
sa
p
r
p
c0
ssa
pr
_ _
kvp0j
kvpsj
; s > s
sa
_
_
_
18
where p
c0
is the yield stress at zero suction, p
r
a reference stress,
and j the elastic compression index. The specic shape of this func-
tion depends on the variation of the compressibility with suction,
i.e. the function k
vp
(s). The suction-increase yield surface can also
be derived from Eq. (5). If the shrinkability (k
vs
) is assumed to be
independent of stress, the suction-increase yield surface is simply:
s s
0
19
where s
0
is the yield suction. Eq. (19) represents a horizontal line in
the stresssuction space. Because Eq. (5) is not dened at zero suc-
tion and zero mean stress, the apparent tensile strength surface
cannot be derived from (18). A separate function is usually
introduced:
p
0
ks 20
where k was assumed to be constant in Alonso et al. [2], but vary
with suction in Georgiadis et al. [38].
If Approach B, i.e. Eq. (7), is adopted to describe the volume
change, the loading-collapse surface can then be written as:
p
0
c

p
0
c0
; s 6 s
sa
p
0
r
p
0
c0
p
r
_ _
k0j
ksj
; s > s
sa
_
_
_
21
The suction-increase surface is the same as Eq. (19). The apparent
tensile strength surface is recovered from (21) by setting p
c0
0:
p
0
0
0 22
which represents the vertical line going through the origin of the
stress space.
If the SFG model, i.e. Eqs. (13) and (15), is adopted to describe
the volume change, the yield stress takes the following form:
p
c

p
c0
s; s 6 s
sa
p
c0
s
sa
s
sa
ln
s
ssa
; s > s
sa
_
23
The function denes the suction-increase surface and is indepen-
dent of the specic function f(s) used in (13). The apparent tensile
strength function is also dened in (23) by setting p
c0
0. The load-
ing-collapse surface takes a very similar form:
p
c

p
cn0
s; s 6 s
sa
p
cn0
p
c0
p
c0
f s s
sa
s
sa
ln
s
ssa
_ _
f s; s > s
sa
_
24
where p
cn0
is the new yield stress at zero suction (Fig. 11). This func-
tion is however dependent on the specic forms of f(s) used in (13).
The functions dened in (18)(24) are all continuous over the
entire ranges of possible suction or pore pressure values. However,
functions (18) and (21) may not be smooth, dependent on func-
tions k
vp
(s) and k(s), respectively. On the other hand, functions
(23) and (24) are continuous and smooth. All these functions can
be incorporated into existing constitutive models for saturated
soils. For example, if the modied Cam clay model is used for sat-
urated soil behaviour, the yield function can be generalised to
unsaturated states along the suction axis:
f q
2
M
2
p p
0
p
c
p 0 25
where f is the yield function in the stress space, q is the deviator
stress, M is the slope of the critical state line in q p space, and
p
0
and p
c
are dened above. Again, Eq. (25) is valid for all pore pres-
sure and suction values.
4.2. Reconstituted soils versus compacted soils
A soil can become partially saturated with water in different
ways. Two types of unsaturated soil samples are often used in lab-
oratory: (1) dry soil powders are statically or dynamically com-
pacted at specied water contents, (2) samples reconstituted
from slurry are dried to unsaturated states. The rst type of sam-
ples (compacted soils) is far more common than the second type
of samples (reconstituted soils), because it is more difcult to
desaturate a slurry sample than a compacted sample. It has been
noted that most constitutive models for unsaturated soils are
based on experimental data for compacted soils [108]. Compacted
soil samples can be prepared at an initial water content smaller
than the optimum water content (compacted dry of optimum) or
larger than the optimum water content (compacted wet of opti-
mum). Reconstituted samples can be air-dried, heat-dried,
freeze-dried or osmosis-dried. Different samples preparation
methods usually result in different soil microstructures. For
example, soils compacted dry of optimum usually have a double-
porosity microstructure, meaning that the pore size distribution
curve exhibits two or more peaks. In these soils, there are two
types of pores: large inter-aggregates pores which are collapsible
upon wetting and small intra-aggregates pore which are more
stable. On the other hand, soils air-dried from slurry usually
exhibit a uni-modal pore size distribution, at least at low stresses.
Nevertheless, as recently pointed out by Tarantino [133], the
boundary between compacted and reconstituted soils is not always
clear and the microstructure of a soil can change with stress and
hydraulic paths.
Unsaturated soils with a bi-modal pore size distribution (PSD)
are usually collapsible. Wetting such a soil can collapse the inter-
aggregates pores and results in a uni-modal pore size distribution
when the soil becomes full saturated. In terms of constitutive mod-
elling, a collapsible soil is characterised by a loading-collapse yield
s
p
Suction increase
yield surface
Apparent tensile
strength
Loading collapse
yield surface
A
B
D
45
o
45
o
45
o
C
c0
p
cn0
p
Fig. 11. Yield stress variation with suction in the SFG model [108].
766 D. Sheng / Computers and Geotechnics 38 (2011) 757776
surface where the yield stress increases with increasing suction [2].
This loading-collapse (LC) yield surface evolves with stress or suc-
tion changes. As showed in Fig. 12, the soil at point A is unsatu-
rated and has a bi-modal PSD and that at point B is saturated
and has a uni-modal PSD. The wetting path A ?B causes the LC
curve to evolve from LC
A
to LC
B
, as the inter-aggregates pores
collapse and the soil volume decreases. However, drying the
uni-modal soil at point B to point A and then compressing it to
point C, i.e. stress path B ?A ?C, should regenerate the bi-modal
PSD, because the soil at point C becomes collapsible again accord-
ing to the BBM [2]. Some experimental data also support such a
development of soil collapsibility [129]. In other words, the pore
size distribution can evolve with stress and hydraulic paths. Most
constitutive models in the literature are based on data for com-
pacted soils and hence predict an evolution of the loading collapse
yield surface as shown in Fig. 12.
The question is now about the reconstituted soils. A soil recon-
stituted from slurry is characterised by a uni-modal PSD, wetting it
under constant stress usually does not cause volume collapse.
However, can such a soil become collapsible if it is air-dried and
compressed to high stresses? According to the SFG model [108],
the yield surface for a slurry soil can evolve into a loading-collapse
surface if the soil is dried and compressed to sufciently high stres-
ses (point C in Fig. 13), which means that a uni-modal reconsti-
tuted soil can evolve into a bi-modal collapsible soil (stress path
B ?A ?C). Is there experimental evidence for such an evolution?
In other words, can drying and compressing of a reconstituted soil
lead to aggregation or cluster formation so that the soil becomes
collapsible? Unfortunately there is very few data on reconstituted
soils in the literature. Nevertheless, the classic reference by Jen-
nings and Burland [48] seems to support such an evolution.
Fig. 14 is a replot from Jennings and Burland [48] for an air-dry silt.
Bi-modal: A
B: Uni-modal
s
p
C: Bi-modal?
LC
A
LC
B
Fig. 12. Evolution of the yield surface for compacted soils according to Alonso et al.
[2].
Slurry: B
s
p
C: Bi-modal?
LC
C
LC
A
LC
B
A
Fig. 13. Evolution of the yield surface for reconstituted soil according to Sheng et al.
[108].
Fig. 14. Oedometer curves for air-dry silt soaked at various constant applied pressures [48].
D. Sheng / Computers and Geotechnics 38 (2011) 757776 767
It is clear that the air-dry soil can become collapsible if it is com-
pressed to sufciently high stresses. Another set of data reported
by Cunningham et al. [20] for a reconstituted silty clay also seem
to indicate that compressing a soil at sufciently high suction
can eventually lead to a collapsible soil (Fig. 15). The limited data
in the literature seem to support the evolution of LC surfaces as
suggested in the SFG model (Fig. 13). However, experimental data
on reconstituted soils are generally too few to be conclusive.
In summary, the microstructure and particularly the pore size
distribution of a soil is usually reected in the yield surface and
volume change equation of the constitutive model for the soil. A
bi-modal pore size distribution can evolve to a uni-modal pore size
distribution under proper stress and hydraulic paths, and vice
versa.
5. Shear strength
The change of shear strength with suction or saturation is one of
the main reasons behind rainfall-induced landslides. It is related to
the volume change equation [110]. However, this relationship has
been overlooked in most existing models for unsaturated soils. If
the slope of the critical state line is assumed to be independent
of suction, such as supported by experimental data of Ng and
Chiu [79], Thu et al. [135] and Nuth and Laloui [83], the shear
strengthsuction relationship can be derived uniquely from the
volume change equation. If the slope of the critical state line
depends on suction, as supported by data of Toll [136], Toll and
Ong [137] and Merchn et al. [74], two equations are needed to
dene the shear strengthsuction relationship, namely the volume
change equation, and the M(s) function, with M being the slope of
the critical state line in the deviator mean stress space.
Bishop and Blight [8] rst proposed an effective stress denition
to interpret the shear strength of unsaturated soils:
s c
0
r
0
n
tan/
0
c
0
r
n
vs tan/
0
c r
n
tan/
0
26
where s is the shear strength, c
0
is the effective cohesion for satu-
rated states and is usually assumed to be zero, r
0
n
and r
n
are respec-
tively the effective and net normal stress on the failure plane, u
0
is
the effective friction angle of the soil, v is the well-known Bishops
effective stress parameter, and c is the apparent cohesion which in-
cludes the friction term due to suction, i.e. c c
0
vs tan/
0
.
Fredlund et al. [25] proposed the following relationship which
conveniently separates the shear strength due to stress from that
due to suction:
s c
0
r
n
u
a
tan/
0
u
a
u
w
tan/
b

c r
n
u
a
tan/
0
27
where s is the shear strength, c
0
the effective cohesion for saturated
states and is usually assumed to be zero, r
n
the normal stress on the
failure plane, u
0
the effective friction angle of the soil, and u
b
the
frictional angle due to suction. Obviously, if u
b
is set to u
0
in Eq.
(27), the Coulomb friction criterion in terms of effective stress for
saturated soils is recovered.
A common conception is that the shear strength of an unsatu-
rated soil can be sufciently dened by a single effective stress
[3,105,146]. However, this conception is only true when the slope
of the critical state line or the friction angle of the soil does not
change with suction. When the friction angle of the soil changes
with suction, the shear strength can no longer be dened by a sin-
gle effective stress. This is clear by comparing equation (26) with
(27):
s c
0
r
n
tan/
0
s s tan/
b

c
0
r
n
s
tan/
b
tan/
0
s
_ _
tan/
0
s c
0
r
0
n
tan/
0
s 28
where the Bishop effective stress parameter (v) is set to tan u
b
/
tan u
0
. It is clear that the variable (s) cannot be eliminated no mat-
ter how the effective stress is dened.
The shear strength of an unsaturated soil is fully dened if the
apparent cohesion (c) or the friction angle due to suction (u
b
),
and the friction angle due to stress u
0
in Eq. (27) are known. A
number of alternative equations have been used in the literature
to dene c or u
b
(e.g. [27,54,75,84,137,140]). Most of these equa-
tions are empirically based and are dened independently from
the volume change equation, which, if incorporated into a constitu-
tive model, may lead to inconsistency with the yield surface as dis-
cussed in the Section above. Rojas [94,95] presented an analytical
expression for Bishops parameter v in terms of the saturated frac-
tion and S
r
of the unsaturated fraction of the soil and used it to
interpret shear strength. One issue with this approach is that v
cannot be easily determined either experimentally or theoretically.
A similar approach was proposed by Pereira and Alonso [90] where
v is expressed in terms of the effective degree of saturation that
depends on the micro- and macro-saturation.
Sheng et al. [115] recently compared various shear strength
equations for unsaturated soils against a large number of data sets.
The equations include those empirically based and those embed-
ded in constitutive models. The equations studied by Sheng et al.
[115] are listed in Table 1. Fig. 16 presents one example of such
comparisons. The parameters used in the shear strength equations
Fig. 15. Isotropic compression curves for a reconstituted silty clay at various
suctions [20].
Table 1
Shear strength equations and parameters used for compacted kaolin clay.
Equations tan u
b
/ tan u
0
Parameters
1. berg and Sllfors [84] S
r
SWCC
2. Fredlund et al. [27] (S
r
)
j
j = 1.3, h
s
= 0.659, SWCC
3. Vanapalli et al. [140] hhr
hs hr
_ _
h
s
= 0.659, SWCC
4. Toll and Ong [137]
Sr Sr2
Sr1Sr2
_ _
k
k = 1.2, SWCC
5. Alonso et al. [2] a a = 0.4
6. Sun et al. [122]
a
sa
a = 110 kPa
7. Khalili and Khabbaz [54] (s
ae
/s)
r
s
ae
= 60 kPa, r = 0.55
8. Sheng et al. [108] ssa
s

ssa
s
_ _
ln
s
ssa
_ _
s
sa
= 25 kPa
768 D. Sheng / Computers and Geotechnics 38 (2011) 757776
are given in Table 1. The comparative study by Sheng et al. [115]
reveals that:
1. If the friction angle of the soil is assumed to be independent of
suction, all shear strength equations in the literature can be for-
mulated in form of either Eq. (26) or Eq. (27). In this case, there
is little difference in formulating shear strength in one single
stress variable or in two independent stress variables. The real
challenge is to nd a single effective stress when the friction
angle depends on suction.
2. The performance of the shear strength equations in predicting
experimental data depends on the careful determination of
material parameters and also on the specic data set. A shear
strength equation may predict one data set better than other
data sets.
3. The shear strength equations that incorporate the soilwater
characteristic curve (SWCC) generally require more parameters.
This group of equations seem to provide reasonable prediction
of shear strength for unsaturated soils (e.g. [27,140,137]). How-
ever, some equations are sensitive to the residual suction
[140,137], which can be difcult to determine accurately from
the SWCC.
4. Simpler shear strength equations that are embedded in consti-
tutive models appear to provide reasonable predictions of shear
strength [54,108,122]. These equations contain typically one or
two parameters. However, these equations usually do not pre-
dict any peak value of the shear strength attained at an interme-
diate suction level.
6. Water retention behaviour and hydro-mechanical coupling
The issue of interaction between the mechanical and hydraulic
behaviour was perhaps rst raised by Wheeler [145] and then by
Dangla et al. [22]. The rst complete model that deals with coupled
hydro-mechanical behaviour of unsaturated soils was perhaps due
to Vaunat et al. [143]. A number of coupled models soon followed
(e.g. [114,149]). With respect to hydraulic behaviour of unsatu-
rated soils, many models [28,139] take advantage of the fact that
the inuence of suction on degree of saturation is more signicant
than the inuence of deformation. The dependency of degree of
saturation on suction is described by a soilwater characteristic
curve (also called soilwater retention curve, SWRC). Only until
recently, the effects of deformation on SWCCs have been consid-
ered (e.g. [31,76,124,149,160]).
As pointed out by Wheeler et al. [149], the mechanical behav-
iour of an unsaturated soil depends on degree of saturation even
if the suction, net stress and specic volume are kept the same
for the soil. Separate treatment of mechanical and hydraulic com-
ponents in modelling unsaturated soil behaviour has certain limi-
tations in reproducing experimental observation. It would be
difcult to consider the saturation dependency in a mechanical
model that is independent of the hydraulic behaviour. Similarly,
a hydraulic model that is independent of mechanical behaviour
cannot easily take into account the effects of soil density on the
SWCC. Experimental data generally demonstrate the following
points:
1. A SWCC obtained under a higher net mean stress tends to shift
towards the higher suction [31,62,73,81]. This means that the
incremental relationship between degree of saturation (S
r
)
and suction (s) depends on net mean stress ( p) or soil density.
2. When the suction is kept constant, isotropic loading or unload-
ing can also change the degree of saturation of an unsaturated
soil [149]. This implies that the degree of saturation is related
to stress or soil density when the suction is kept constant.
One of the early models that fully couple the hydraulic and
mechanical components of unsaturated soil behaviour is that by
Wheeler et al. [149]. Models that appeared before or soon after
Wheeler et al. [149] tend to accentuate the inuences of the
hydraulic component on the mechanical component, not vice versa
(e.g. [83,114,143]). The interaction between the mechanical and
hydraulic components in the model by Wheeler et al. [149] was
realised through the use of the average soil skeleton stress (effec-
tive stress), the modied suction and the coupling between the
loading-collapse and suction-increase and suction-decrease sur-
faces. The average soil skeleton stress r

ij
is an amalgam of stress,
suction and degree of saturation. The modied suction (s

) is a
combination of suction and porosity. Therefore, the inuence of
hydraulic behaviour on the stressstrain relationship is considered
via the denition of the average stress. The inuence of porosity on
the saturationsuction relationship is considered via the denition
of the modied suction. The model by Wheeler et al. [149] is one of
the few models that are qualitatively tenable in terms of coupling
mechanical behaviour with hydraulic behaviour for unsaturated
600
500
400
300
200
0 100 200 300 400 0 100 200 300 400
600
500
400
300
200
D
e
v
i
a
t
o
r

S
t
r
e
s
s
,

k
P
a

Suction, kPa Suction, kPa
Predictions
Test data (after Thu et al. 2007b)
1
Confining pressure: 100 kPa
7
6
5
3
2
4
8
Test data (after Thu et al. 2007b)
Confining pressure: 200 kPa
Predictions
7
8
5
1
2
4
3
6
Fig. 16. Predictions of triaxial test data on compacted kaolin clay by various shear strength equations (data by [135]). Note that the prediction by Khalili and Khabbaz [54], i.e.
No. [7], can be improved signicantly by using a lower air entry value.
D. Sheng / Computers and Geotechnics 38 (2011) 757776 769
soils. However, the use of the modied suction and the soil skele-
ton stress, which is one of the advantages that makes the model
rigorously consistent in thermodynamics, can become a disadvan-
tage as well, particularly in terms of quantitative prediction and
the application of the model. For example, one of the difculties
in using this model is to quantify the synchronised movement be-
tween the loading-collapse (LC) surface and the suction-increase
(SI) and suction-decrease (SD) surfaces. This synchronicity cannot
easily be calibrated by laboratory experiments [93] or dened
theoretically.
In more recent models, the inuences of mechanical properties
on the hydraulic behaviour are usually modelled via the depen-
dency of the SWCC on soil volume [31,132], soil density [71,124],
or volumetric strain [82]. Gallipoli et al. [31] suggested including
a function of specic volume (v) in the SWCC equation of van
Genuchten [139]. Tarantino [132] showed there is a unique rela-
tionship between the water ratio (product of S
r
and e) and the mat-
ric suction and used this relationship to modify van Genuchtens
equation. The modied van Genuchtens equation takes a similar
form as Gallipoli et al. [31]. It is also common to express the SWCC
equation in incremental forms. For example, Sun et al. [124] pro-
posed a hydraulic model in the following form:
dS
r
k
se
de k
ss
ds=s 29
where k
ss
is the slope of main drying or wetting curve, and k
se
the
slope of degree of saturation versus void ratio curve under constant
suction. In theory, k
ss
in the equation above can only be determined
from constant-volume tests (de = 0). Man [71] used a similar
equation as (29). In his model both air entry value (s
ae
) and the
slope of main drying curve (k
ss
) vary with void ratio. Nuth and
Laloui [82] provided an alternative approach of modelling SWCC
for a deforming soil. They assume there is an intrinsic SWCC for a
non-deforming soil and deformation of the soil can shift this intrin-
sic SWCC along the suction axis. The shift is governed by an air en-
try value that depends on the volumetric strain.
The models by Sun et al. [124], Nuth and Laloui [82], Man [71]
and many others (e.g. [53,83,108,114,128,160]) essentially all
adopt a water retention equation in the following form:
dS
r
. . .ds . . .de
v
30
This equation is not wrong, but the embedded S
r
s relationship is
for constant volume (de
v
= 0). Therefore, it does not recover the con-
ventional SWCC equations, which are obtained under constant
stress. The volume change along a conventional SWCC can be signif-
icant, particularly at low suctions. For expansive soils, it is also com-
mon to study the water retention behaviour under constant volume
[65,96,98]. On the other hand, if one specic S
r
s relation is valid for
constant stress, the S
r
s relation for constant volume would inevita-
bly involve soil compressibility and hence stress (suction) level and
stress (suction) history. Therefore, it is unlikely that the same S
r
s
relation can be used to calibrate water retention data both for con-
stant stress and constant volume. Furthermore, neglecting the vol-
ume change along SWCC can lead to inconsistent prediction of the
degree of saturation during undrained compression, an issue raised
by Zhang and Lytton [154].
Sheng and Zhou [116] proposed a new method for coupling
hydraulic with mechanical behaviour. This new method is based
on the fact that SWCCs are obtained under constant stress. In
Sheng and Zhou [116], the volume change behaviour and the water
retention behaviour under isotropic stress states are represented
by the following incremental equations, respectively:
de
v
Ad p Bds 31
dS
r
Eds
S
r
n
1 S
r

f
Ad p
E B
S
r
n
1 S
r

f
_ _
ds
S
r
e
1 S
r

f
de 32
where parameters A and B are related to the specic volume change
equation as discussed in Section 3, parameter E is the gradient of
the conventional SWCC, e is the void ratio, n is the porosity, and f
is a tting parameter. If the SFG model is used to describe the vol-
ume change, parameters A and B would then take the following
form respectively:
A
k
vp
p f s
; B
k
vs
p f s
33
The S
r
s relationship in Eq. (32) is dened for constant stress
(d p 0) and hence parameter E refers to the gradient of the conven-
tional SWCC:
E
d S
SWCC
r
s
_ _
ds
34
where S
SWCC
r
represents the conventional SWCC equation. The void
ratio in Eq. (32) refers to the value at the current stress and suction.
It is clear from Eq. (32) that the S
r
s relationship for constant
volume (de = 0) is more complex than the conventional SWCC
equation (dS
r
Eds). Eq. (32) was proposed based on experimental
observation as well as the intrinsic phase relationship for un-
drained condition:
dS
r

S
r
n
de
v
; dw 0 35
A
B
LC yield surface
SI surface
SD surface
s
p
C
D
A
B
Main drying curve
Main wetting curve
s
S
r
C
D
A
B
Main drying curve
Main wetting curve
s
S
r
C
D
Scanning curve
(a) stress path in net stress-suction
space
(b) inconsistent change of S
r
(c) consistent change of S
r
A
Fig. 17. Qualitative analysis of isotropic compression under undrained condition.
770 D. Sheng / Computers and Geotechnics 38 (2011) 757776
where w is the gravimetric water content. Eq. (35) actually imposes
a constraint on suction change under undrained compression. This
constraint is obtained by substituting Eq. (35) into Eq. (32):
S
r
S
r
1 S
r

f
Ad p nE Bds; dw 0 36
Zhang and Lytton [154] recently noted that some models fall short
in predicting undrained behaviour of unsaturated soils. The specic
issue raised is illustrated in Fig. 17. Assume the initial state of a soil
is inside the elastic zone, but on the main wetting curve, i.e. point A
in Fig. 17a. Compressing the soil under undrained condition will
lead to some suction decrease [127,131], say to point B. Assume B
is still inside the elastic zone. Unloading from B to A will recover
the initial volume of the soil and hence the initial degree of satura-
tion should be recovered as well. However, if the volume change
along the SWCC is neglected, the change of S
r
along path ACB would
follow the main wetting curve and is hence elastoplastic, whereas
the change of S
r
along path BDA would follow the scanning curve
and is hence elastic, leading to inconsistent change of S
r
(Fig. 17b). It would then seem unlikely that a model where the irre-
versible volume change is not synchronised with the irreversible
saturation change could lead to a consistent prediction of saturation
change over the closed path ABA [155]. This inconsistency is actu-
ally due to the assumption that the main wetting curve is dened
for constant volume.
If Eq. (32) is used, the inconsistence in Fig. 17b will be avoided.
Because of Eq. (35), the model will predict no change of S
r
as long
as e
v
= 0, irrespective of evolution of the suction-decrease surface.
Eq. (35) is satised as long as the suction changes according to
(36). It is easy to understand that the loading path ACB is not on
the initial main wetting curve and the unloading path BDA is not
on the scanning curve, because the mean stress is not constant
along those stress paths. The main wetting curve at point A also
shifts to that at point B, as the mean stress changes (Fig. 17c).
The synchronised evolution of the LC, SI and SD surfaces (as in
[149] is not necessary for the consistent prediction in Fig. 17c. In-
deed, the suction path can be elastoplastic even though the stress
path is elastic.
The following equation can be derived from Eq. (32):
@S
r
@e

S
r
1 S
r

f
e
; ds 0 37
The void ratio (e) in Eq. (37) refers to the initial void ratio at the cur-
rent stress and its variation is purely due to stress change. It can
also be interpreted as the initial void ratio at the start of the SWCC
tests. Eq. (37) shows that the SWCC for a soil shifts with its initial
void ratio. This is similar to the approach by Gallipoli et al. [31]
where the van Genuchten equation was modied to include the ini-
tial void ratio. Indeed, their SWCC equation can be re-written as:
@S
r
@e
mnw
S
r
1 S
1=m
r

e
38
where m and n are two tting parameters in the original van
Genuchten equation, and w is another parameter introduced by
Gallipoli et al. [31]. If the product (mnw) was set to 1, Eq. (38) would
be equivalent to (37). Sheng and Zhou [116] showed that the intrin-
sic phase relationship requires:
1 S
r
e
P
@S
r
@e
P
S
r
e
39
The above constraint is satised if (mnw) = 1 in Eq. (38). It is also
interesting to note that a value of 1.1 for (mnw was used in all
the numerical examples in Gallipoli et al. [31].
Eq. (37) can be integrated either analytically for certain f values
or numerically in more general cases. Because Eq. (37) is in an
incremental form, integration of the equation requires one specic
SWCC equation that corresponds to a reference initial void ratio. In
other words, the conventional SWCC equation is only used for the
reference initial void ratio and the new SWCC for a new initial void
ratio is obtained by integration of (37).
The model by Sheng and Zhou [116] is validated against a vari-
ety of data sets. In Fig. 18, the results from suction-controlled
oedometer compression tests by Jotisankasa [51] are used to vali-
date Eq. (37). Since the suction remains constant for each curve in
Fig. 18, no SWCC equation is required here and the only parameter
required is f. Eq. (37) can directly be used to predict the variation
of degree of saturation due the variation of stress-induced volume
change. In this case, the void ratio (e) in Eq. (37) is the current void
ratio and it changes due to stress variation in the oedometer tests.
Fig. 18 clearly shows that Eq. (37) predicts very well the relation-
ship between the degree of saturation and the void ratio when
the suction is kept constant. Indeed, the experimental data are al-
most exactly on the predicted S
r
e curves. In Fig. 19, the data by
Vanapalli et al. [141] on a compacted till were used. The SWCC
for e
0
= 0.517 was tted by the van Genuchten equation, while
the other SWCCs are predicted by Eq. (37) with f = 0.03. Fig. 19
shows that both the slope and the air entry value of the SWCCs
change with the initial void ratio. In the two cases studied, the
model by Sheng and Zhou [116] seems to be able to capture the ef-
fect of initial void ratio on the soil water retention behaviour. One
0.30
0.20
0.10
0.00
S
r
7-10-U
Calibration: 5-10-K
7-10-V
7-10-P
7-10-W 5-10-L
5-10-I
Predictions
= 0.1
1.2 1.4 1.6 1.8
1+e
Fig. 18. Validation of Eq. (37) using suction-controlled oedometer compression
data of a compacted soil (data after [51]).
0.6
0.7
0.8
0.9
1
1.1
0.1 1 10
s (kPa)
100 1000 10000
e
0
=0.444
e
0
=0.474
e
0
=0.514
e
0
=0.517
Predictions
S
r

Fig. 19. Predicted and measured SWCCs at various initial void ratios of a compacted
till (data after [141]).
D. Sheng / Computers and Geotechnics 38 (2011) 757776 771
advantage of this approach is that Eq. (37) can be used with any
existing SWCC equations, including those for uni-modal pore size
distribution (e.g. [28,139]) and those for bi-modal pore size distri-
bution (e.g. [11,40]), as well as those advanced models that adopt
hysteretic scanning loops within the main drying-wetting loop
[64,8789].
7. Finite element implementation
One of the ultimate goals of constitutive modelling is to imple-
ment the model in a numerical method to solve boundary value
problems. A constitutive model can generally be formulated in
the following incremental form [66,110,114]:
d r
dh
_ _

D
ep
W
ep
T H
_ _
de
ds
_ _
40
where r is the stress vector, e is the vector of soil skeleton strain, h is
the volumetric water content, D
ep
, W
ep
, T and H are constitutive
matrices [110].
In the displacement nite element method, the nodal displace-
ments and pore pressures are rst solved from the equilibrium and
continuity equations, based on the current stress states and the
current volumetric water content. The strain and suction incre-
ments are then derived from the displacements and pore pres-
sures. For given strain and suction increments, the current stress
vector, the volumetric water content and the internal variables
must be updated according to Eq. (40). This update is generally car-
ried out by numerical integration.
One of the main challenges in integrating Eq. (40) arises from
the non-convexity of the yield surface around the transition be-
tween saturated and unsaturated states [106,113,147]. The non-
convexity seems to persist irrespective of the stress variables and
is demonstrated in Fig. 20 [111].
Both implicit and explicit schemes have been used to integrate
unsaturated soils models. In implicit schemes, all gradients and
functions are estimated at advanced unknown stress states and
the solution is achieved by iteration. This group of stress integra-
tion schemes include Vaunat et al. [142], Borja [10], Hoyos and
Arduino [46], Zhang and Zhou [153], and Tamagnini and De Genna-
ro [130]. They usually do not deal with the non-convex problem by
avoiding the transition between saturated and unsaturated states.
Borja [10] was perhaps the only one in this group who noticed the
problem. He suggested keeping the step size sufciently small to
avoid overshooting the plastic zone by the elastic trial stress
(Fig. 20). Otherwise, there is currently no general method used to
tackle the non-convexity problem in implicit schemes.
On the other hand, explicit schemes estimate the gradients and
functions at the current known stress states and proceed in an
incremental fashion. These methods are theoretically more
appropriate for non-convex models [87,88]. Gonzlez and Gens
[42] compared both an implicit and an explicit scheme for integrat-
ing the BBM and found that the latter is more robust and efcient.
However, explicit schemes usually require to determine the inter-
section between the current yield surface and the elastic trial
stress path and some substepping methods to control the integra-
tion error [102,110,111,112,113,119,120].
A key issue in integrating the incremental stressstrain rela-
tionships using an explicit scheme is thus to nd the intersection
between the elastic trial stress path and the current yield surface.
The most complex situation occurs when the yield surface is
crossed more than once, such as shown in Fig. 20. However, it
is not possible to know a priori how many times the yield surface
is crossed. Therefore, for non-convex yield surfaces, the key
task is to nd the very rst intersection for any possible stress
and hydraulic path.
Finding the intersection between the elastic trial stress incre-
ment and the current yield surface can be cast as a problem of nd-
ing the multiple roots of a nonlinear equation:
f a f r
a
; s
a
; z
a
0 41
where 0 6 a 6 1; f r; s; z is the yield function, z is a set of internal
variables, and subscript a indicates the quantity is estimated at
strain increment aDe and suction increment aDs. Pedroso et al.
[87,88] proposed a novel method to bracket the roots (a). The
method is illustrated in Fig. 21. For a given increment (a = 1), the
number of roots of f(a) is rst computed. If there is more than
one root, the increment is divided into two equal sub-increments.
The number of roots of each sub-increment is then computed. If
the rst sub-increment contains more than one root, it is further
divided into two sub-increments. This process is repeated until
the rst sub-increment contains at most one root (Fig. 21). Once
the roots are bracketed, the solution of the rst root can be found
by using numerical methods such as the Pegasus method [118].
Sheng et al. [111] applied the root bracketing method by Pedr-
oso et al. [87,88] to integrate the SFG model and found that the
method can indeed provide an accurate solution of the intersection
45
o
p
s
Elastic zone
p
s
Elastic zone
unsaturated
saturated
p
saturated
Elastic trial stress
Elastic zone
unsaturated
s
45
o
(a) net stress
(Alonso et al.,1990)
(b) effective stress
(Sheng et al., 2004)
(c) net stress
(Sheng et al., 2008a)
unsaturated
saturated
LC surface
Fig. 20. Non-convexity of yield surfaces in unsaturated soil models in the suctionstress space.
f
0

1
1/2
1/4
1/8
Fig. 21. Bracketing the roots for nonlinear function according to Pedroso et al.
[87,88].
772 D. Sheng / Computers and Geotechnics 38 (2011) 757776
problem. However, this method is found to be computationally
expensive. The formula used to nd the number of roots of a non-
linear function requires both the rst and second orders of gradi-
ents of the function and requires numerical integration.
Furthermore, the root-nding procedure must be applied for all
suction increments near the non-convexity. Further research in
this direction is clearly required.
8. Conclusions
Some conclusions can be drawn from this review of constitutive
modelling of unsaturated soils:
1. Partial saturation is only a state of soil. Constitutive models for
soils should represent the soil behaviour over entire ranges of
possible stress and pore pressure values. This requires a consis-
tent and coherent merge of fundamental soil mechanics princi-
ples for both saturated and unsaturated states.
2. The volume change behaviour is one of the most fundamental
properties of soils. For soils in unsaturated states, the volume
change equation also underpins the yield stresssuction and
shear strengthsuction relationships. It also affects the soil
water retention behaviour. Indeed, one of the most essential
differences amongst various existing models is the volume
change equation. Other differences are often the consequences
of the volume change equation.
3. Three groups of volumetric models are compared and it is
shown that each has advantages and disadvantages. There is
certainly room for improvement of all these models. One obser-
vation here is that it seems difcult to describe the volume
change of unsaturated soils in terms of one single stress
variable.
4. The methods used to dene the loading-collapse, suction-
increase and apparent tensile strength surfaces should be
consistent with the volume change equation. In addition, these
surfaces and the shear strength function are all related to each
other and hence should be dened consistently with each other.
5. It seems possible to model reconstituted soils and compacted
soils within the same theoretical framework. The pore size dis-
tribution evolves with mechanical and hydraulic loading and is
reected by the loading collapse yield surface and the volume
change equation.
6. If the friction angle of the soil is assumed to be independent of
suction, all shear strength equations in the literature can be for-
mulated either in terms of one single stress variable or in terms
of two independent stress variables. The real challenge is to nd
a single effective stress when the friction angle depends on
suction.
7. The performance of the shear strength equations in predicting
experimental data depends on the careful determination of
material parameters and also on the specic data set. A shear
strength equation may predict one data set better than other
data sets.
8. When coupling the hydraulic component with the mechanical
component in a constitutive model, it is recommended to take
into account the volume change along soilwater characteristic
curves. Neglecting this volume change can lead to inconsistent
prediction of volume and saturation changes.
9. Unsaturated soil models are characterised by non-convex yield
surfaces at the transition between saturated and unsaturated
states. This non-convexity, if handled rigorously, can signi-
cantly complicate the implementation of these models into
nite element codes. In this case, an explicit stress integration
scheme incorporating an efcient root nding algorithm is
preferred.
Acknowledgements
The author would like thank the editor of Computers and Geo-
technics for the invitation of this review paper. This paper is based
on the General Report presented at the 5th International Conference
on Unsaturated Soils held at Barcelona during 68 September 2010
[105]. A number of people provided valuable comments and con-
tributions to the General Report: Y.J. Cui, D.G. Fredlund, D. Gallipol-
i, A. Gens, S.L. Houston, J. Kodikara, J.M. Pereira, W.T. Solowski, D.A.
Sun, C. Yang, X. Zhang and A.N. Zhou. The author is grateful to all of
them. The General Report has since been substantially revised to
produce this paper, partly inspired by questions and informal dis-
cussions during and after the Barcelona conference. Finally the
nancial support from the Australian Research Council is greatly
appreciated.
References
[1] Al-Badran Y, Schanz T. Yielding surface model of volume change
characteristics for unsaturated ne grained soils. In: Buzzi O, Fityus SG,
Sheng D, editors. Unsaturated soils theoretical & numerical advances in
unsaturated soil mechanics. CRC Press; 2009. p. 86371.
[2] Alonso EE, Gens A, Josa A. A constitutive model for partially saturated soils.
Gotechnique 1990;40:40530.
[3] Alonso EE, Pereira JM, Vaunat J, Olivella S. A microstructurally based effective
stress for unsaturated soils. Geotechnique 2010;60:91325.
[4] Alonso EE, Vaunat J, Gens A. Modelling the mechanical behaviour of expansive
clays. Eng Geol 1999;54:17383.
[5] Arson C, Gatmiri B. On damage modelling in unsaturated clay rocks. Phys
Chem Earth 2008;33:63540.
[6] Baker R, Frydman S. Unsaturated soil mechanics, critical review of physical
foundations. Eng Geol 2009;106:2639.
[7] Bian H, Shahrour I. Numerical model for unsaturated sandy soils under cyclic
loading: application to liquefaction. Soil Dynam Earthq Eng 2009;29:23744.
[8] Bishop AW, Blight GE. Some aspects of effective stress in saturated and partly
saturated soils. Gotechnique 1963;13:17797.
[9] Bolzon G, Schreer BA, Zienkiewicz OC. Elastoplastic soil constitutive laws
generalised to partially saturated states. Gotechnique 1996;46:27989.
[10] Borja RI. Cam clay plasticity, Part V: a mathematical framework for three
phase deformation and strain localization analysis of partially saturated
porous media. Comp Methods Appl Mech Eng 2004;193:530138.
[11] Burger CA, Shackelford CD. Evaluating dual porosity of pelletized
diatomaceous earth using bimodal soil-water characteristic curve functions.
Can Geotech J 2001;38:5366.
[12] Buscarnera G, Nova R. An elastoplastic strainhardening model for soil
allowing for hydraulic bonding-debonding effects. Int J Numer Anal
Methods Geomech 2009;33:105586.
[13] Cardoso R, Alonso EE. Degradation of compacted marls: a microstructural
investigation. Soils Found 2009;49:31528.
[14] Chiu CF, Ng CWW. A state-dependent elasto-plastic model for saturated and
unsaturated soils. Gotechnique 2003;53(9):809929.
[15] Cleall PJ, Seetharam SC, Thomas HR. Inclusion of some aspects of chemical
behaviour of unsaturated soil in thermo/hydro/chemical/mechanical models.
I: Model development. ASCE J Eng Mech 2007;133:33847.
[16] Coussy O, Pereira JM, Vaunat J. Revisiting the thermodynamics of hardening
plasticity for unsaturated soils. Comput Geotech 2010;37:20715.
[17] Cui YJ, Delage P. Yielding and plastic behaviour of an unsaturated compacted
silt. Gotechnique 1996;46(2):291311.
[18] Cui YJ, Sultan N, Delage P. A thermomechanical model for clays. Can Geotech J
2000;37:60720.
[19] Cui YJ, Sun DA. Constitutive modelling: from isothermal to non-isothermal
behaviour of unsaturated soils. In: Buzzi O, Fityus SG, Sheng D, editors.
Unsaturated soils theoretical & numerical advances in unsaturated soil
mechanics. CRC Press; 2009. p. 493506.
[20] Cunningham MR, Ridley AM, Dineen K, Burland JB. The mechanical behaviour
of a reconstituted unsaturated silty clay. Gotechnique 2003;53:18394.
[21] Dafalias YF. Bounding surface plasticity. I: mathematical foundation and
hypoelasticity. ASCE J Eng Mech 1986;112:96687.
[22] Dangla P, Malinsky L, Coussy O. Plasticity and imbibition-drainage curves for
unsaturated soils: a unied approach. In: Numog VI, Pietruszczak S, Pande
GN, editors. Numerical models in geomechanics. Rotterdam: Balkema; 1997.
p. 1416.
[23] Delage P, Graham J. Understanding the behaviour of unsaturated soils
requires reliable conceptual models: state of the art report. In: Alonso EE,
Delage P, editors. Unsaturated soils, vol. 3. Rotterdam: Balkema; 1996. p.
122356.
[24] DOnza F, Gallipoli D, Wheeler SJ. Effect of anisotropy on the prediction of
unsaturated soil response under triaxial and oedometric conditions. In:
Alonso EE, Gens A, editors. Unsaturated soils, vol. 2. CRC Press; 2011. p.
78794.
D. Sheng / Computers and Geotechnics 38 (2011) 757776 773
[25] Fredlund DG, Morgenstern NR, Widger A. Shear strength of unsaturated soils.
Can Geotech J 1978;15:31321.
[26] Fredlund DG, Rahardjo H. Soil mechanics for unsaturated soils. New
York: John Wiley & Sons; 1993.
[27] Fredlund DG, Xing A, Fredlund MD, Barbour SL. Relationship of the
unsaturated soil shear strength to the soil-water characteristic curve. Can
Geotech J 1996;33:4408.
[28] Fredlund DG, Xing A. Equations for the soil-water characteristic curve. Can
Geotech J 1994;31:52132.
[29] Gallipoli D, Gens A, Chen G, DOnza F. Modelling unsaturated soil behaviour
during normal consolidation and at critical state. Comput Geotech
2008;35:82534.
[30] Gallipoli D, Gens A, Sharma R, Vaunat J. An elastoplastic model for
unsaturated soil incorporating the effects of suction and degree of
saturation on mechanical behaviour. Gotechnique 2003;53:12335.
[31] Gallipoli D, Wheeler SJ, Karstunen M. Modelling of variation of degree of
saturation in a deformable unsaturated soil. Gotechnique 2003;53:10512.
[32] Gens A. Constitutive modelling: application to compacted soils. In: Alonso EE,
Delage P, editors. Unsaturated soils, vol. 3. Rotterdam: Balkema; 1996. p.
1179200.
[33] Gens A. Some issues in constitutive modelling of unsaturated soils. In: Buzzi
O, Fityus SG, Sheng D, editors. Unsaturated soils theoretical & numerical
advances in unsaturated soil mechanics. CRC Press; 2009. p. 61326.
[34] Gens A. Soil-environmental interactions in geotechnical engineering.
Gotechnique 2010;60:374.
[35] Gens A, Alonso EE. A framework for the behaviour of unsaturated expansive
clays. Can Geotech J 1992;29:101332.
[36] Gens A, Snchez M, Sheng D. On constitutive modelling of unsaturated soils.
Acta Geotech 2006;1:13747.
[37] Gens A, Guimars LdoN, Snchez M, Sheng D. Developments in modelling the
generalised behaviour of unsaturated soils. In: Toll DG, Augarde CE, Gallipoli
D, Wheeler SJ, editors. Unsaturated soils: advances in geo-engineering. CRC
Press; 2008. p. 5362.
[38] Georgiadis K, Potts DM, Zdravkovic L. Three-dimensional constitutive model
for partially and fully saturated soils. Int J Geomech 2005;5:24455.
[39] Gili JA, Alonso EE. Microstructural deformation mechanisms of unsaturated
granular soils. Int J Numer Anal Methods Geomech 2002;26:43368.
[40] Gitirana GDFN, Fredlund DG. Soil-water characteristic curve equation with
independent properties. J Geotech Geoenviron Eng 2004;130:20912.
[41] Gray WG, Schreer BA. Thermodynamic approach to effective stress in
partially saturated porous media. Eur J Mech A/Solids 2001;20:52138.
[42] Gonzlez NA, Gens A. Evaluation of a constitutive model for unsaturated
soils: stress variable and numerical implementation. In: Alonso EE, Gens A,
editors. Unsaturated soils, vol. 2. CRC Press; 2011. p. 82936.
[43] Guimars LdoN, Gens A, Olivella S. Coupled termo-hydro-mechanical and
chemical analysis expansive clay subjected to heating and hydration.
Transport Porous Media 2007;66:34172.
[44] Houlsby GT. The work input to an unsaturated granular material.
Gotechnique 1997;47(1):1936.
[45] Houston SL. Applied unsaturated soil mechanics, state of the art report. In:
Juca JFT, de Campos TMP, Marinho FAM, editors. Unsaturated soils, vol.
3. Lisse: Balkema; 2002. p. 112734.
[46] Hoyos LR, Arduino P. Implicit algorithms in modeling unsaturated soil
response in three-invariant stress space. Int J Geomech 2008;8:26673.
[47] Hutter K, Laloui L, Vulliet L. Thermodynamically based mixture models for
saturated and unsaturated soils. Mech Cohes Frict Mater 1999;4:295338.
[48] Jennings JEB, Burland JB. Limitations to the use of effective stresses in partly
saturated soils. Gotechnique 1962;12:12544.
[49] Jiang MJ, Leroueil S, Konrad JM. Insight into shear strength functions of
unsaturated granulates by DEM analyses. Comput Geotech 2004;31:47389.
[50] Jommi C. Remarks on the constitutive modelling of unsaturated soils. In:
Tarantino A, Mancuso C, editors. Experimental evidence & theoretical
approaches in unsaturated soils. Rotterdam: Balkema; 2000. p. 13953.
[51] Jotisankasa A. 2005. Collapse behaviour of a compacted silty clay, Ph.D.
Thesis. Imperial College London, London.
[52] Katti DR, Schmidt SR, Ghosh P, Katti KS. Molecular modelling of the
mechanical behaviour and interactions in dry and slightly hydrated sodium
montmorillonite interlayer. Can Geotech J 2007;44:42535.
[53] Khalili N, Habte MA, Zargarbashi S. A fully coupled ow deformation model
for cyclic analysis of unsaturated soils including hydraulic and mechanical
hystereses. Comput Geotech 2008;35:87289.
[54] Khalili N, Khabbaz MH. A unique relationship for the determination of the
shear strength of unsaturated soils. Gotechnique 1998;48:6817.
[55] Kikumoto M, Kyokawa H, Nakai T, Shahin HM. A simple elastoplastic model
for unsaturated soils and interpretations of collapse and compaction
behaviours. In: Alonso EE, Gens A, editors. Unsaturated soils, vol. 2. CRC
Press; 2011. p. 84955.
[56] Kimoto S, Oka F, Fushita T, Fujiwaki M. A chemo-thermo-mechanically
coupled numerical simulation of the subsurface ground deformations due to
methane hydrate dissociation. Comput Geotech 2007;34:21628.
[57] Kohgo Y, Nakano M, Miyazaki T. Theoretical aspects of constitutive modelling
for unsaturated soils. Soils Found 1993;33:4963.
[58] Kohgo Y. Review of constitutive models for unsaturated soils and initial-
boundary value analyses. In: Karube D, Iizuka A, Kato S, Kawai K, Tateyama K,
editors. Unsaturated soils geotechnical and geoenvironmental issues, proc 2
Asian conf; 2003. p. 2140.
[59] Kohler R, Hofstetter G. A cap model for partially saturated soils. Int J Numer
Anal Methods Geomech. 2008;32:9811004.
[60] Kolymbas D. An outline of hypoplasticity. Arch Appl Mech 1991;61:
14351.
[61] Kurucuk N, Kodikara J, Fredlund DG. Evolution of the compaction process:
experimental study preliminary results. In: Buzzi O, Fityus SG, Sheng D,
editors. Unsaturated soils theoretical & numerical advances in unsaturated
soil mechanics. CRC Press; 2009. p. 88793.
[62] Lee IM, Sung SG, Cho GC. Effect of stress state on the unsaturated shear
strength of a weathered granite. Can Geotech J 2005;42:62431.
[63] Li XS. Thermodynamics-based constitutive framework for unsaturated soils,
1: Theory. Gotechnique 2007;57:41122.
[64] Li XS. Modelling of hysteresis response for arbitrary wetting/drying paths.
Comput Geotech 2005;32:1337.
[65] Lloret A, Villar MV, Snchez M, Gens A, Pintado X, Alonso EE. Mechanical
behaviour of heavily compacted bentonite under high suction changes.
Gotechnique 53, 2740.
[66] Lloret M, Snchez M, Wheeler SJ. Generalised elastoplastic stressstrain and
modied suction-degree of saturation relations of fully coupled model. In:
Buzzi O, Fityus SG, Sheng D, editors. Unsaturated soils theoretical &
numerical advances in unsaturated soil mechanics. CRC Press; 2009. p.
66772.
[67] Loret B, Hueckel T, Gajo A. Chemo-mechanical coupling in saturated porous
media: elastic-plastic behaviour of homoionic expansive clays. Int J Solids
Struct 2002;39:2773806.
[68] Loret B, Khalili N. An effective stress elasticplastic model for unsaturated
porous media. Mech Mater 2002;34:97116.
[69] Manzanal D, Pastor M, Fernndez Merodo JA. A unied constitutive model for
unsaturated soils based on generalised plasticity theory. In: Buzzi O, Fityus
SG, Sheng D, editors. Unsaturated soils theoretical & numerical advances in
unsaturated soil mechanics. CRC Press; 2009. p. 65560.
[70] Marshall TJ, Holmes JW, Rose CW. Soil physics. New York: Cambridge
University Press; 1996.
[71] Man D. Predicting the dependency of a degree of saturation on void ratio
and suction using effective stress principle for unsaturated soils. Int J Numer
Anal Methods Geomech 2010;34:7390.
[72] Man D, Khalili N. A hypoplastic model for mechanical response of
unsaturated soils. Int J Numer Anal Methods Geomech 2008;32:190326.
[73] Matyas EL, Radhakrishna HS. Volume change characteristics of partly
saturated soils. Gotechnique 1968;18:43248.
[74] Merchn V, Vaunat J, Romero E, Meca T. Experimental study of the inuence
of suction on the residual friction angle of clays. In: Toll DG, Augarde CE,
Gallipoli D, Wheeler SJ, editors. Unsaturated soils: advances in geo-
engineering. CRC Press. p. 4238.
[75] Miao L, Houston SL, Cui Y, Yuan J. Relationship between soil structure
mechanical behaviour for an expansive unsaturated clay. Can Geotech J
2007;44:12637.
[76] Miller GA, Khoury CN, Muraleetharan KK, Liu C, Kibbey TCG. Effects of soil
skeleton deformations on hysteretic soil water characteristic curves:
experiments and simulations. Water Resour Res 2008;44:W00C06.
[77] Modaressi A, Modaressi H. Thermoplastic constitutive model for
unsaturated soils: a prospective approach. In: Pande GN, Pietruszczak S,
editors. Numerical models in geomechanics. Rotterdam: Balkema; 1995. p.
4550.
[78] Morvan M, Wong H, Branque D. An unsaturated soil model with minimal
number of parameters based on bounding surface plasticity. Int J Numer Anal
Methods Geomech 2010;34:151237.
[79] Ng CWW, Chiu ACF. Behaviour of loosely compacted unsaturated volcanic
soil. J Geotech Geoenviron Eng 2001;127:102736.
[80] Ng CWW, Cui YJ, Chen R, Delage P. The axis-translation and osmotic
techniques in shear testing of unsaturated soils. Soils Found 2007;47:67584.
[81] Ng CWW, Pang YW. Inuence of stress state on soil-water characteristics and
slope stability. J Geotech Geoenviron Eng 2000;126:15766.
[82] Nuth M, Laloui L. Advances in modelling hysteretic water retention curve in
deformable soils. Comput Geotech 2008;35:83544.
[83] Nuth M, Laloui L. Effective stress concept in unsaturated soils: clarication
and validation of a unied framework. Int J Numer Anal Methods Geomech
2008;32:771801.
[84] berg A, Sllfors G. Determination of shear strength parameters of
unsaturated silts and sands based on the water retention curve. Geotech
Test J 1997;20:408.
[85] Oldecop LA, Alonso EE. Theoretical investigation of the time-dependent
behaviour of rockll. Gotechnique 2007;57:289301.
[86] Pastor M, Zienkiewicz O, Chan A. Generalized plasticity and the modelling of
soil behaviour. Int J Numer Anal Methods Geomech 1990;14:15190.
[87] Pedroso DM, Sheng D, Sloan SW. Stress update algorithm for elastoplastic
models with non-convex yield surfaces. Int J Numer Methods Eng
2008;76:202962.
[88] Pedroso DM, Sheng D, Zhao J. The concept of reference curves for constitutive
modelling in soil mechanics. Comput Geotech 2008;36:14965.
[89] Pedroso DP, Williams DJ. A novel approach for modelling soil-water
characteristic curves with hysteresis. Comput Geotech 2010;37:37480.
[90] Pereira JM, Alonso EE. Insights into the links between microstructure and
Bishops v parameter. In: Buzzi O, Fityus SG, Sheng D, editors. Unsaturated
soils theoretical & numerical advances in unsaturated soil mechanics. CRC
Press; 2009. p. 68590.
774 D. Sheng / Computers and Geotechnics 38 (2011) 757776
[91] Pereira JM, de Gennaro V. On the time-dependent behaviour of unsaturated
geomaterials. In: Alonso EE, Gens A, editors. Unsaturated soils, vol. 2. CRC
Press; 2011. p. 9215.
[92] Pereira JM, Wong H, Dubujet P, Dangla P. Adaptation of existing behaviour
models to unsaturated states: Application to CJS model. Int J Numer Anal
Methods Geomech 2005;29:112755.
[93] Raveendiraraj A. Coupling of mechanical behaviour and water retention
behaviour in unsaturated soils. PhD thesis. University of Glasgow, UK; 2009.
[94] Rojas E. Equivalent stress equation for unsaturated soils, I: Equivalent stress.
Int J Geomech 2008;8:28590.
[95] Rojas E. Equivalent stress equation for unsaturated soils, II: Solid porous
model. Int J Geomech 2008;8:2919.
[96] Romero E, Gens A, Lloret A. Water permeability, water retention and
microstructure of unsaturated compacted Boom clay. Eng Geol
1999;54:11727.
[97] Romero E, Jommi C. An insight into the role of hydraulic history on the
volume changes of anisotropic clayey soil. Water Resour Res
2008;44:W12412.
[98] Romero E, Gens A, Lloret A. Suction effect on a compacted clay under iso-
thermal conditions. Gotechnique 2003;53:6581.
[99] Russell AR, Khalili N. A unied bounding surface plasticity model for
unsaturated soils. Int J Numer Anal Methods Geomech 2006;30:181212.
[100] Samat S, Vaunat J, Gens A. A thermomechanical framework for modelling the
response of unsaturated soils. In: Toll DG, Augarde CE, Gallipoli D, Wheeler SJ,
editors. Unsaturated soils: advances in geo-engineering. CRC Press; 2008. p.
54752.
[101] Snchez M, Gens A, Guimares LdoN, Olivella S. A double structure
generalized plasticity model for expansive materials. Int J Numer Anal
Methods Geomech 2005;29:75187.
[102] Snchez M, Gens A, Guimaraes LdoN, Olivella S. Implementation algorihm of
a generalised plasticity model for swelling clay. Comput Geotech
2008;35:86071.
[103] Santagiuliana R, Schreer BA. Enhancing the Bolzon-Schreer-Zienkiewicz
constitutive model for partially saturated soil. Transport Porous Media
2006;65:130.
[104] Scholts L, Hicher PY, Nicot F, Chareye B, Darve F. On the capillary stress
tensor in wet granular materials. Int J Numer Anal Methods Geomech
2009;33:1289313.
[105] Sheng D. Constitutive modelling of unsaturated soils: discussion of
fundamental principles. In: Alonso EE, Gens A, editors. Unsaturated soils,
vol. 1. CRC Press; 2011. p. 91112.
[106] Sheng D. Non-convexity of the Barcelona Basic Model comment on SJ
Wheeler, D Gallipoli & M Karstunen (2002; 26:1561-1571). Int J Numer Anal
Methods Geomech 2003;27:87981.
[107] Sheng D, Fredlund DG. Elastoplastic modelling of unsaturated soils: an
overview. Keynote lecture, 12th Int conf int assoc computer methods &
advances in geomechanics (IACMAG), 16 October, 2008, Goa, India, Indian
Institute of Technology, Bombay; 2008. p. 2084105.
[108] Sheng D, Fredlund DG, Gens A. A new modelling approach for unsaturated
soils using independent stress variables. Can Geotech J 2008;45:51134.
[109] Sheng D, Fredlund DG, Gens A. Reply to the discussion by Zhang and Lytton
on A new modelling approach for unsaturated soils using independent stress
variables. Can Geotech J 2008;45:178894.
[110] Sheng D, Gens A, Fredlund DG, Sloan SW. Unsaturated soils: from constitutive
modelling to numerical algorithms. Comput Geotech 2008;35:81024.
[111] Sheng D, Pedroso DM, Abbo AJ. Stress path dependency and non-convexity of
unsaturated soil models. Comput Mech 2008;42:68595.
[112] Sheng D, Sloan SW, Gens A, Smith DW. Finite element formulation and
algorithms for unsaturated soils. Part I: Theory. Int J Numer Anal Methods
Geomech 2003;27:74565.
[113] Sheng D, Smith DW, Sloan SW, Gens A. Finite element formulation and
algorithms for unsaturated soils. Part II: Verication and application. Int J
Numer Anal Methods Geomech 2003;27:76790.
[114] Sheng D, Sloan SW, Gens A. A constitutive model for unsaturated soils:
thermomechanical andcomputational aspects. Comput Mech2004;33:45365.
[115] Sheng D, Zhou AN, Fredlund DG. Shear strength criteria for unsaturated soils,
Geotech Geol Eng, vol. 29; in press. doi:10.1007/s10706-009-927-x.
[116] Sheng D, Zhou AN. Coupling hydraulic with mechanical models for
unsaturated soils. Can Geotech J 2011;48:82640.
[117] Sivakumar V, Wheeler SJ. Inuence of compaction procedure on the
mechanical behaviour of an unsaturated compacted clay. Part 1: Wetting
and isotropic compression. Gotechnique 2000;50:35968.
[118] Sloan SW, Abbo AJ, Sheng D. Rened explicit integration of elastoplastic
models with automatic error control. Engineering Computations, 18, 121-54.
Erratum: Eng Comput 2001;19:594594.
[119] Soowski WT, Gallipoli D. Explicit stress integration with error control for the
Barcelona Basic Model. Part I: Algorithms formulation. Comput Geotech
2009;37:5967.
[120] Soowski WT, Gallipoli D. Explicit stress integration with error control for the
Barcelona Basic Model. Part II: Algorithms efciency and accuracy. Comput
Geotech 2009;37:6881.
[121] Stropeit K, Wheeler SJ, Cui YJ. An anisotropic elasto-plastic model for
unsaturated soils. In: Toll DG, Augarde CE, Gallipoli D, Wheeler SJ, editors.
Unsaturated soils: advances in geo-engineering. CRC Press; 2008. p. 25632.
[122] Sun DA, Matsuoka H, Yao YP, Ichihara W. An elastoplastic model for
unsaturated soil in three-dimensional stresses. Soils Found 2000;40:1728.
[123] Sun DA, Cui HB, Matsuoka H, Sheng D. A three-dimensional elastoplastic
model for unsaturated compacted soils with hydraulic hysteresis. Soils Found
2007;47:25364.
[124] Sun DA, Sheng D, Cui HB, Sloan SW. A density-dependent elastoplastic hydro-
mechanical model for unsaturated compacted soils. Int J Numer Anal
Methods Geomech 2007;31:125779.
[125] Sun DA, Sheng D, Sloan SW. Elastoplastic modelling of hydraulic and stress
strain behaviour of unsaturated soils. Mech Mater 2007;39:21221.
[126] Sun DA, Sheng D, Xu XF. Collapse behaviour of unsaturated compacted soil.
Can Geotech J 2007;44:67386.
[127] Sun DA, Sheng D, Xiang L, Sloan SW. Elastoplastic prediction of hydro-
mechanical behaviour of unsaturated soils under undrained conditions.
Comput Geotech 2008;35:84552.
[128] Sun DA, Sun W, Xiang L. Effect of degree of saturation on mechanical
behaviour of unsaturated soils and its elastoplastic simulation. Comput
Geotech 2010;37:67888.
[129] Suriol J, Gens A, Alonso EE. Behavior of compacted soils in suction-controlled
oedometer. In: Proc 2nd int confunsat soils. Beijing: International Academic
Publishers; 1998. p. 43843.
[130] Tamagnini R, De Gennaro V. Implicit integration of extended Cam-clay model
for unsaturated soils. In: Toll DG, Augarde CE, Gallipoli D, Wheeler SJ, editors.
Unsaturated soils: advances in geo-engineering. CRC Press; 2008. p. 7139.
[131] Tang AM, Cui YJ, Barnel N. Compression-induced suction change in a
compacted expansive clay. In: Toll DG, Augarde CE, Gallipoli D, Wheeler SJ,
editors. Unsaturated soils: ad vances in geo-engineering. CRC Press; 2008. p.
36974.
[132] Tarantino A. A water retention model for deformable soils. Gotechnique
2009;59:75162.
[133] Tarantino A. Unsaturated soils: compacted with reconstituted states. In:
Alonso EE, Gens A, editors. Unsaturated soils, vol. 1. CRC Press; 2011. p.
11336.
[134] Thu TM, Rahardjo H, Leong EC. Elastoplastic model for unsaturated soil with
incorporation of the soil-water characteristic curve. Can Geotech J
2007;44:6777.
[135] Thu TM, Rahardjo H, Leong EC. Critical state behaviour of a compacted silt
specimen. Soils Found 2007;47:74955.
[136] Toll DG. A framework for unsaturated soil behaviour. Gotechnique
1990;40:3144.
[137] Toll DG, Ong BH. Critical state parameters for an unsaturated residual sandy
clay. Gotechnique 2003;53:93103.
[138] Unno T, Kazama M, Uzuoka R, Sento N. Liquefaction of unsaturated sand
considering the pore air pressure and volume compressibility of the soil
particle skeleton. Soils Found 2008;48:87100.
[139] van Genuchten MT. A closed-form equation for predicting the hydraulic
conductivity of unsaturated soils. Soil Science Society of America Journal
1980;44:8928.
[140] Vanapalli SK, Fredlund DG, Pufahl DE, Clifton AW. Model for the prediction
of shear strength with respect to soil suction. Can Geotech J 1996;33:
37992.
[141] Vanapalli SK, Fredlund DG, Pufahl DE. The inuence of soil structure and
stress history on soil-water characteristics of a compacted till. Gotechnique
1999;49:14359.
[142] Vaunat J, Cante JC, Ledesma A, Gens A. A stress point algorithm for an
elastoplastic model in unsaturated soils. Int J Plasticity 2000;16:12141.
[143] Vaunat J, Romero E, Jommi C. An elastoplastic hydromechanical model for
unsaturated soils. In: Tarantino A, Mancuso C, editors. Experimental evidence
& theoretical approaches in unsaturated soils. Rotterdam: Balkema; 2000. p.
12138.
[144] Vlahinic I, Jennings HM, Andrade JE, Thomas JJ. A novel and general form of
effective stress in a partially saturated porous material: the inuence of
microstructure. Mech Mater 2011;43:2535.
[145] Wheeler SJ. Inclusion of specic water volume within an elasto-plastic model
for unsaturated soil. Can Geotech J 1996;33:4257.
[146] Wheeler SJ. Constitutive modelling of unsaturated soils. Keynote lecture
presented at 4th international conference on unsaturated soil (unpublished
PDF presentation), Arizona, USA; 2006.
[147] Wheeler SJ, Gallipoli D, Karstunen M. Comments on use of the Barcelona
Basic Model for unsaturated soils. Int J Numer Anal Methods Geomech
2002;26:156171.
[148] Wheeler SJ, Karube D. Constitutive modelling. In: Alonso EE, Delage P, editors.
Unsaturated soils, vol. 3. Rotterdam: Balkema; 1996. p. 132356.
[149] Wheeler SJ, Sharma RS, Buisson MSR. Coupling of hydraulic hysteresis and
stressstrain behaviour in unsaturated soils. Gotechnique 2003;53:4154.
[150] Wheeler SJ, Sivakumar V. An elasto-plastic critical state framework for
unsaturated soil. Gotechnique 1995;45:3553.
[151] Yang C, Cui YJ, Pereira JM, Huang MS. A constitutive model for unsaturated
cemented soils under cyclic loading. Comput Geotech 2008;35:8539.
[152] Zhang F, Ikariya T. A new model for unsaturated soil using skeleton stress and
degree of saturation as state variables. Soils Found 2011;51:6781.
[153] Zhang HW, Zhou L. Implicit integration of chemo-plastic constitutive mode
for partially saturated soils. Int J Numer Anal Methods Geomech
2008;32:171535.
[154] Zhang X, Lytton RL. Discussion on A new modelling approach for unsaturated
soils using independent stress variables. Can Geotech J 2008;45:17847.
[155] Zhang X, Lytton RL. Modied state-surface approach to unsaturated soil
behaviour study. Part I: Basic concept. Can Geotech J 2009;46:53652.
D. Sheng / Computers and Geotechnics 38 (2011) 757776 775
[156] Zhang X, Lytton RL. Modied state-surface approach to unsaturated soil
behaviour study. Part II: General formulation. Can Geotech J 2009;46:55370.
[157] Zhao CG, Liu Y, Gao FP. Work and energy equations and the principle of
generalised effective stress for unsaturated soils. Int J Numer Anal Methods
Geomech 2010;34:92036.
[158] Zhou AN, Sheng D. Yield stress, volume change and shear strength behaviour
of unsaturated soils: Validation of the SFG model. Can Geotech J
2009;46:103445.
[159] Zhou AN, Sheng D, Sloan SW, Gens A. Interpretation of unsaturated soil
behaviour in the stress saturation space. Comput Geotech, submitted for
publication.
[160] Zhou C. Modelling of suction effect on fabric yielding and kinematic
hardening of reconstituted soils. In: Buzzi O, Fityus SG, Sheng D, editors.
Unsaturated soils theoretical & numerical advances in unsaturated soil
mechanics. CRC Press; 2009. p. 62934.
776 D. Sheng / Computers and Geotechnics 38 (2011) 757776

Você também pode gostar