You are on page 1of 24

1249

Challenges to modelling heave in expansive soils


Hung Q. Vu and Delwyn G. Fredlund

Abstract: There are challenges associated with the numerical modelling of unsaturated expansive soils. The challenges
are primarily related to the quantification of the void ratio constitutive surface, the characterization of the void ratio
constitutive surface at low stresses and (or) suction, and the solution of coupled equations with several nonlinear unsaturated soil property functions. This study suggests that the void ratio constitutive surface of an expansive soil subject
to a monotonic wetting path can be estimated from volume change indices obtained from conventional laboratory tests.
The constitutive surfaces for both the soil structure and the water phase can be described using mathematical equations
that allow net normal stress and suction to be reduced to zero. The solutions for two typical volume change problems
are presented using both a coupled approach and an uncoupled approach. The first example problem simulates water
leakage from a pipe under a flexible cover. The second example problem simulates the infiltration of water at ground
surface. The results of the analyses are in accordance with anticipated behaviour. The results also show that the answers from an uncoupled analysis compared well with those from a coupled analysis. It is suggested that an uncoupled
analysis may be adequate for most prediction of heave problems involving unsaturated expansive soils.
Key words: heave prediction, numerical modelling, expansive soil, constitutive surface, uncoupled analysis, matric suction.
Rsum : Il y a des dfis associs avec la modlisation numrique de sols non saturs gonflants. Les dfis sont principalement relis la quantification de la surface constitutive du rapport de vides, la caractrisation de la surface constitutive de rapport des vides faibles contraintes et (ou) succion, et la solution des quations couples avec plusieurs
fonctions non linaires des proprits des sols non saturs. Cette tude suggre que la surface constitutive du rapport
des vides dun sol gonflant soumis un cheminement de mouillage monotonique peut tre estime partir des indices
de changement de volume obtenus par des essais conventionnels en laboratoire. Les surfaces constitutives tant pour la
structure du sol que pour la phase deau peuvent tre dcrites au moyen dquations mathmatiques qui permettent de
rduire zro la contrainte normale et la succion nettes. On prsente les solutions pour deux problmes typiques de
changement de volume au moyen tant dapproche couple que non couple. Le premier exemple de problme simule la
perte deau dun tuyau sous un couvert flexible. Le second exemple simule linfiltration de leau la surface du terrain. Les rsultats des analyses sont en accord avec le comportement anticip. Les rsultats montrent galement que les
rponses de lanalyse non couple se comparent bien avec ceux de lanalyse couple. Lon suggre que lanalyse non
couple peut tre adquate pour la plupart des problmes de prdiction de soulvement impliquant des sols gonflants
non saturs.
Mots cls : prdiction de soulvement, modlisation numrique, sol gonflant, surface constitutive, analyse non couple,
matrice de succion.
[Traduit par la Rdaction]

Vu and Fredlund

1272

Introduction
The behaviour of an unsaturated expansive soil can be formulated using the theory of unsaturated soils, formulated using two independent stress state variables, the constitutive
relationships for the soil structure and water phase, and flow
laws for the water phase (Alonso et al. 1990, Gens and
Alonso 1992, Fredlund and Rahardjo 1993, Vu and Fredlund
Received 29 September 2005. Accepted 15 May 2006.
Published on the NRC Research Press Web site at
http://cgj.nrc.ca on 5 January 2007.
H.Q. Vu.1,2 Clifton Associates Ltd., 340 Maxwell Crescent,
Regina, SK S4N 5Y5, Canada.
D.G. Fredlund. Golder Associates Ltd., 145 1st Ave N.,
Saskatoon, SK S7K 1W6, Canada.
1
2

Corresponding author (e-mail: hung_vu@golder.com).


Present address: Golder Associates Ltd., 145 1st Ave. N.,
Saskatoon, SK S7K 1W6, Canada

Can. Geotech. J. 43: 12491272 (2006)

2004). The two stress state variables are net normal stress,
( ua), and matric suction (ua uw), where is total normal
stress, ua is pore-air pressure, and uw is pore-water pressure.
Changes in the void ratio and degree of saturation of an unsaturated soil can be expressed as functions of the stress state
variables to form two three-dimensional constitutive surfaces
(Matyas and Radhakrishna 1968).
The elasticity parameter functions required for a volume
change analysis can be computed from the constitutive surfaces. Testing techniques are available for directly measuring
the soil properties associated with the constitutive surfaces
(Fredlund and Rahardjo 1993; Pham et al. 2004); however,
the tests are costly, laborious, and time consuming and may
not be justifiable for routine engineering projects. Approximate volume change (and water content) coefficients should
be adequate for most engineering analyses. Procedures for
the approximation of the volume change coefficients assist
in the implementation of unsaturated soil mechanics into
geotechnical engineering practice (Fredlund 2000a).

doi:10.1139/T06-073

2006 NRC Canada

1250

The primary interactive processes involved with a volume


change analysis of an unsaturated expansive soil are stress
deformation and water flow. The assumption is made that
isothermal conditions exist and the air phase is continuous
and at atmospheric pressure. The stress-deformation process
is governed by static equilibrium, while the water flow is
governed by water continuity equation. Solutions for volume
change require that both the equilibrium equation and the
continuity equation be solved. The solutions can be obtained
using either a coupled or an uncoupled approach.
Coupled solutions are difficult to obtain in part because of
the nonlinear soil property functions associated with both
water flow and stress deformation. An uncoupled solution is
obtained by independently solving the water flow process
and the stress-deformation process. Uncoupled solutions can
be more easily achieved than coupled solutions because the
soil property functions involved in each process (i.e., water
flow or stress deformation) are considered to be independent
of one another. Coupling between the stress deformation and
the water flow process takes place under transient conditions. Under steady-state conditions, the coupling disappears
and pore-water pressure conditions can be determined
through the solutions of the continuity equation.
The objective of this paper is to investigate difficulties
that have been encountered when modelling heave in expansive soils when using unsaturated soil theory. The specific
objectives of this paper involve the: (i) estimation of the void
ratio constitutive surface of an unsaturated expansive soil
from swelling indices obtained from conventional oedometer
tests; (ii) review of existing constitutive surface equations
for both the soil structure and the water phase; (iii) presentation and solution of problems associated with low net normal stresses and (or) low matric suctions when an expansive
soil swells; (iv) examination and evaluation of proposed
equations for the constitutive surfaces; and (v) presentation
and comparison of uncoupled and coupled solutions for two
typical volume change problems using the constitutive surfaces suggested in this study.

Governing partial differential equations


Vu and Fredlund (2004) presented the formulations of
governing partial differential equations for both saturated
unsaturated seepage and stress deformation. The equations
were derived based on the following assumptions: (i) the air
phase is continuous and at atmospheric pressure; (ii) the soil
is isotropic and nonlinear elastic (i.e., incrementally elastic);
(iii) strains are small; (iv) the pore water is incompressible;
and (v) the effects of air diffusing through water, air dissolving in the water, and the movement of water vapour are negligible. The incremental elasticity equations have a form
similar to those presented by Biot (1941) for a soil with occluded air bubbles.
Let us consider a two-dimensional field with x and y as
the rectangular Cartesian coordinates (i.e., x for the horizontal direction and y for the vertical direction) and with u and v
as components of the displacement vector (i.e., u and v for
the x- and y-directions, respectively). The governing partial
differential equation for water flow through a heterogeneous,
anisotropic, saturatedunsaturated soil can be derived by satisfying conservation of mass for a representative elemental

Can. Geotech. J. Vol. 43, 2006

volume, REV, assuming that water flow follows Darcys law


with a nonlinear hydraulic conductivity (Vu and Fredlund
2004):
[1]

w1

v
(ua uw)
x uw

Y
+ w2
=
+
kw

t
t
x x wg

y u

w + Y
kw

y wg

where kwx and kwy are the hydraulic conductivities in the xand y-directions, respectively; t is time; w is the density of
water, g is the gravitational acceleration, Y is the elevation
head, and
w1 =

m1w
E
, or
s
m1
(1 2) Ew

w2 = m 2w

1
3E
m1wm 2s
, or

Hw (1 2)HEw
m1s

where E is an elasticity parameter for the soil structure with


respect to a change in the net normal stress; H is an elasticity parameter for the soil structure with respect to a change
in matric suction; Ew is an elasticity parameter for the
change in the amount of water in the soil with respect to a
change in the net normal stress; Hw is an elasticity parameter
for the change in the amount of water in the soil with respect
to a change in matric suction; is Poissons ratio; coefficients of volume change, m1s and m 2s, are slopes of the soil
structure constitutive surface; and coefficients of volume
changes, m1w and m 2w, are slopes on the constitutive surface
for the water phase. The general form for the unloading constitutive surfaces for both the soil structure and the water
phase are shown in Fig. 1.
The governing partial differential equations for soil structure equilibrium, written in terms of displacements in the xand y-directions (i.e., u and v) for the plane strain case of an
isotropic, nonlinear elastic soil are as follows (Vu and
Fredlund 2004):
[2]

[3]

u v
u
v
c
c
c
+
+
+
11
12
33
y y x
x x
y
(ua uw)
ds
+ bx = 0
x
c33

v
u v u
+
+ c22
+
c12

y
x y x y x
(ua uw)
ds
+ by = 0
y

where bx and by are body forces in the x- and y-directions,


respectively; and
c11 = c22 =
c12 =

(1 ) E
(1 + ) (1 2)

E
(1 + ) (1 2 )
2006 NRC Canada

Vu and Fredlund

1251

Fig. 1. Constitutive surfaces for (a) soil structure and (b) water phase of an unsaturated expansive soil.

c33 =
ds =

E
2 (1 + )

[5]

E
(1 2) H

The seepage equation (eq. [1]) shows the influence of the


compressibility and the rate of the volume change of the soil
structure on the transient water flow process in an expansive
soil. Solutions to the seepage equation (eq. [1]) and soil
structure equilibrium equations (eqs. [2] and [3]) can be obtained through the use of either an uncoupled or a coupled
solution. Procedures for uncoupled and coupled analyses are
presented later in this paper.

Soil properties required for volume change


predictions in an expansive soil
Soil properties required for a volume change analysis include: (i) Poissons ratio, ; (ii) an elasticity parameter (i.e.,
elasticity function) for the soil structure with respect to net
normal stress, E; (iii) an elasticity parameter (i.e., elasticity
function) for the soil structure with respect to matric suction,
H; (iv) an elasticity type parameter for the water phase
with respect to net normal stress, Ew; (v) an elasticity type
parameter for the water phase with respect to matric suction,
Hw; and (vi) the hydraulic conductivity function, kw. The soil
properties are functions of both net normal stress and matric
suction. Assuming a value of Poissons ratio, the elasticity
parameters, E, H, Ew, and Hw, can be calculated from the coefficient of volume change, m1s, m 2s, m1w, and m 2w, respectively.
At a particular stress state, the coefficients of volume
change, m1s and m 2s, can be obtained by differentiating the
constitutive surface for the soil structure and written as a
function of void ratio:
[4]

m1s =

dv
d (mean ua )

1
de
1 + e0 d (mean ua )

m 2s =

dv
1
de
=
d (ua uw) 1 + e0 d (ua uw)

where dv = de/(1 + e0), e0 is the initial void ratio of the soil


(i.e., referential element), e is the void ratio of the soil, and mean
is the mean net total stress (i.e., mean = (x + y + z)/3).
The elasticity parameters for the soil structure, E and H,
can be calculated as follows:
[6]

E =3

[7]

H =

(1 2)
m1s

3
m 2s

The coefficients of water volume change, m1w and m 2w, can


be obtained by differentiating the constitutive surface for the
water phase
[8]

m1w =

d w
d (mean ua )

[9]

m 2w =

d w
d (ua uw)

where w = Vw/V0 is the volumetric water content.


The coefficients of water volume change can also be calculated from the void ratio constitutive surface and degree of
saturation, S, constitutive surface as follows:
[10]

m1w =

S
e
de
dS
+
1 + e0 d (mean ua ) 1 + e0 d (mean ua )

[11]

m 2w =

de
dS
S
e
+
1 + e0 d (ua uw) 1 + e0 d (ua uw)

The elasticity parameters for the water phase, Ew and Hw,


can be calculated as follows:
[12]

Ew =

3
m1w
2006 NRC Canada

1252

[13]

Can. Geotech. J. Vol. 43, 2006

Hw =

1
m 2w

It is suggested that the void ratio surface can be estimated


from the swelling indices obtained from extreme planes (i.e.,
net normal stress plane and matric suction plane). The elasticity parameters can then be calculated from the estimated
void ratio constitutive surface. Soil data required to estimate
the void ratio surface are shown in Fig. 2; namely, (i) the
swelling index with respect to net normal stress, Cs; (ii) the
swelling index with respect to matric suction, Cm; (iii) the
initial void ratio, e0, and (iv) the swelling pressure, Ps. The
swelling indices, Cs and Cm, are the slopes of the void ratio
versus logarithm of net normal stress and logarithm of
matric suction, respectively.

Estimation of void ratio constitutive surface


from volume change indices
The following assumptions are made with regard to the
void ratio constitutive surface: (i) there is a linear relationship of void ratio versus logarithm of net normal stress at the
extreme net normal stress plane (i.e., suction equal to s0);
(ii) there is a linear relationship of void ratio versus logarithm of matric suction at an extreme suction plane (i.e., net
normal stress equal to ( ua)0); (iii) the void ratio versus
logarithm of net normal stress and void ratio versus loga-

[14]

rithm of matric suction converge at a net normal stress equal


to ( ua)0 (i.e., to be defined) and a suction equal to s0 (i.e.,
to be defined), and (iv) a constant void ratio plane intersects
the void ratio constitutive surface as a straight line on an
arithmetic plot of the stress state variables.
The assumption of a linear relationship between net normal stress and matric suction at a constant void ratio on an
arithmetic scale is supported by the work of Escario (1969)
(Fig. 3) and Matyas (1969) (Fig. 4). The same lines form asymptotic curves on a logarithmic scale. The shape of these
lines suggests the form for the void ratio surface on a threedimensional plot.
Figure 5 illustrates the first, second, and third assumptions
mentioned above. These assumptions are reasonable and
widely accepted (Holtz and Gibbs 1956; Richards et al.
1984; Ho et al. 1992; Fredlund and Rahardjo 1993).
Swelling pressures, Ps, obtained at different suctions are not
the same; however, at low matric suctions, the difference is
negligible. Figures 5 and 6 are used to derive the void ratio
relationship with respect to net normal stress and matric suction. It should be noted that Fig. 5 is on a semilogarithmic
scale while Fig. 6 is on an arithmetic scale.
The derivation of an equation for the calculation of matric
suction corresponding to any void ratio and net normal
stress, or the calculation of net normal stress from the void
ratio and matric suction is presented in detail in Appendix
A. The equation has the following form:

P Cs / Cm

s0 s
10 ( e e0 )/ Cm 1 [Ps 10 ( e e0 )/ Cs ( ua )]
a

(ua uw) =
+ s0
( e e0 )/ Cs
Ps 10
( ua ) 0

The net normal stress and matric suction in eq. [14] can
be obtained either from a one-dimensional loading test (i.e.,
(y ua) and (ua uw)) or from an isotropic loading test
(i.e., (mean ua) and (ua uw)). While the swelling indices
are essentially the same for both loading conditions (Graham
and Li 1985; Al-Shamrani and Al-Mhaidib 2000; Vu 2003),
appropriate values for the parameters, ( ua)0 and s0, must
be used.
A mathematical equation is needed to describe the entire
void ratio surface for modelling and analysis purposes. This
equation can be used to fit experimental data or data generated from swelling indices (i.e., data generated from
eq. [14]).
Review of equations for the void ratio constitutive
relationship
Mathematical equations have been used to describe the
variation in void ratio or water content with respect to
changes in the stress state of a soil. These mathematical
equations can be divided into two groups. The first group includes those equations that are capable of representing the
shape of the soil property function. These equations are determined for a particular soil under specific testing conditions using a best fit of measured data. These equations
contain fitting parameters that need to be determined. Examples are the equations proposed to best fit the properties of

soils related to the water phase such as the soil-water characteristic curve (SWCC) and the hydraulic conductivity
functions. These equations can be used to describe a wide
range of soil properties. The Fredlund and Xing (1994)
equation, along with a correction factor, can be used to describe the SWCC with soil suctions varying up to a limiting
value of 106 kPa.
The second group includes equations that use other soil
properties as parameters in the equation. Examples include
the equations for shear strength (i.e., using cohesion and angle of shearing resistance) and volume changes (i.e., using
volume change indices). The parameters in these equations
usually have a clear physical meaning; however, the equations might not be suitable to describe the entire range of
stress states. The linear equation on a semilogarithmic plot,
which is commonly used to define the volume change index,
cannot be used at extremely low net total stress conditions
and low soil suctions. These equations need to be modified
to provide greater flexibility to best fit measured data.
Soil properties for an unsaturated soil are known to be
functions of both stress state variables (i.e., net total stress
and matric suction). Sometimes equations are proposed to
relate a soil property to only one dominant stress state variable. These equations can be expanded to be functions of
both stress state variables provided the soil properties can be
evaluated. This section reviews mathematical equations that
2006 NRC Canada

Vu and Fredlund

1253

Fig. 2. Soil property functions required for volume change analysis of an unsaturated expansive soil.

can be used to describe the void ratio constitutive surfaces


associated with an unsaturated swelling soil.
Terzaghi (1943) and Casagrande (1936) noted that the virgin compression curve and the rebound curve of the void ratio versus logarithm of effective stress form essentially a
straight line. The void ratio equation has the following form:
[15]

e = a + b log()

where is the designation of the stress state, and a and b are


constants (i.e., fitting parameters).
Fredlund (1979) suggested that the void ratio constitutive
surface for an unsaturated soil could be linearized over a
wide range of stress changes using the logarithm of the
stress state variables. The void ratio under any set of stress
conditions has the following form:
[16]

e = a + b log( ua ) + c log(ua uw)

where a, b, and c are constants (i.e., fitting parameters).


Lloret and Alonso (1985) studied a number of mathematical equations for the description of the volume change constitutive surface of unsaturated soils subjected to confined or
isotropic compression. The equations were used to best fit
experimental results on different soil types, and the optimum
equations were selected on the basis of minimum fitting errors. For a limited range in the total stress, Lloret and
Alonso (1985) suggested that a suitable analytical expression for the void ratio constitutive surface is as follows:
[17]

e = a + b ( ua ) + c log(ua uw)
+ d ( ua ) log(ua uw)

If the range in the stress variation is large, a more suitable


equation for the void ratio surface was given by (Lloret and
Alonso 1985)
[18]

e = a + b log( ua ) + c log(ua uw)


+ d log( ua ) log(ua uw)

Fredlund (2000b) proposed a four-parameter equation that


can be used to represent an overconsolidated soil that contains both a recompression curve and a virgin compression
branch. The equation of void ratio as a function of stress has
the following form for particular soil suction:
[19]

2
2
e = a + b ln 1 + + d ln 1 +
c
f

where a, b, c, d, and f are constants (i.e., fitting parameters)


at a constant suction.

Low net normal stress and zero suction


problem
A practical engineering problem is encountered in the low
net normal stress and zero suction ranges when the elastic
moduli are calculated from the swelling index, Cs, or other
volume change indices as illustrated in Fig. 7. On the extreme planes, the relationships between void ratio and the
logarithm of net normal stress or matric suction are essentially linear (Fig. 7a). When the semilogarithm relationship
is converted to an arithmetic plot (i.e., void ratio versus net
normal stress or matric suction), the void ratio tends to increase to infinity as the stress state variable approaches zero
(Fig. 7b). Therefore, the calculated elastic modulus with re 2006 NRC Canada

1254

Can. Geotech. J. Vol. 43, 2006

Fig. 3. Cross-sections of void ratio surfaces plotted on arithmetic scale and logarithmic scale for a remoulded Madrid clay (modified from
Escario 1969).

Fig. 4. Matric suction versus isotropic net normal stress at constant void ratio (modified from Matyas 1969).

spect to changes in net normal stress, E, becomes extremely


small (and approaches zero) at low net normal stresses
(Fig. 7c). The elastic modulus with respect to a change in
matric suction, H, also becomes extremely small at low
matric suctions (or zero suction) (Fig. 7c). These unrealistically small values of elastic moduli at low net normal stress
and matric suction result in an unstable solution in numerical modelling. As well, the solutions produce unreasonably
large deformations.
A part of this study was directed at finding a procedure
that could adequately describe void ratio changes at low net
normal stresses and matric suctions. The procedure should
involve the use of an equation in a mathematically continuous form that would adequately describe the void ratio in the
low net normal stress range and low matric suction range.
The equation should also produce a continuous, smooth,
physically reasonable, and differentiable form over the entire
range of net normal stresses and matric suctions.
The equation can then be used to fit a set of measured
data or data calculated from swelling indices. This equation
needs to be able to include a value of void ratio at zero suction and zero net normal stress for modelling and analysis
purposes as illustrated in Fig. 8. Using this approach, unreal 2006 NRC Canada

Vu and Fredlund

1255

Fig. 5. Linear relationships between void ratio and logarithm of stress state variables at extreme planes.

Fig. 6. Relationship between net normal stress and matric suction at constant void ratio plane.

istically large values of void ratio at low net normal stresses


and matric suctions can be avoided.

Proposed equations for void ratio


constitutive surface
Vu (2003) proposed six functions to fit the void ratio constitutive surface of an unsaturated, expansive soil. Two of the
functions; namely unsat-1 with three fitting parameters
(eq. [20]) and unsat-6 with six fitting parameters (eq. [21])
are as follows:
[20]

e = a + b log[1 + ( ua ) + c (ua uw)]

[21]

1 + c ( ua ) + d (ua uw)
e = a + b log

1 + f( ua ) + g (ua uw)

It can be noted that parameter a is the void ratio at zero


net normal stress and matric suction; if this parameter is
known, then the number of fitting parameters is reduced by
one. Parameters c, d, f, and g take on non-negative values to
satisfy the logarithmic equation. A higher number of fitting
parameters will increase the fit of the equation to the data,
but the physical interpretation of the parameters is rapidly
lost (Lloret and Alonso 1985).

Evaluation of equations for the void ratio


constitutive surface
Equations for the void ratio constitutive surfaces were evaluated to determine the most suitable equation for a swelling
soil analysis. Soil data for Regina clay (Shuai 1996) was used
to evaluate these equations for the void ratio surface. The ex 2006 NRC Canada

1256
Fig. 7. Illustration of the problem associated with low net normal stress and zero suction.

Can. Geotech. J. Vol. 43, 2006

eters. The value of R2 closest to 1, with the lowest value of


AIC, indicates the best fit.
Several equations are considered for the best-fit analysis
of the void ratio data generated from the swelling indices;
namely Fredlund (1979); Lloret and Alonso (1985);
Fredlund (2000b); and the equations suggested in this study
(i.e., unsat-1 and unsat-6 functions). The best-fit results
from different void ratio equations to the generated void ratio data (Fig. 9) are presented in Table 1. The Fredlund
(2000) equation has been modified to allow best fit in three
dimensions. Statistical results on the proposed void ratio
equations are presented in Table 2. Figure 10 presents the
comparison of different equations using both the AIC criterion and the R2 criterion. It can be seen that the unsat-6
function appears to provide the best fit to this set of data.
The fitting results using the unsat-6 function are shown in
Fig. 11.
Vu (2003) also evaluated mathematical equations for the
description of the measured void ratio data and measured
water content data for the constitutive surfaces. The results
of the evaluation suggest that the unsat-6 function is satisfactory for describing both the soil structure and water phase
constitutive surfaces.

Example problems using the soil properties


of Regina clay
Two example problems are used to illustrate the suggested
approach to volume change analyses. Experimental data obtained from tests on compacted specimens of Regina clay
are used for the analysis. The void ratio surface for threedimensional unloading conditions is estimated from the
swelling indices (eq. [14]) and described using the unsat-6
equation (eq. [21]). The initial matric suction in the soil
mass is assumed to be constant and equal to 400 kPa for
both examples.

perimental data are obtained under K0-loading conditions. A


dataset for one-dimensional loading conditions was generated, using eq. [14], from the swelling indices for Regina
clay. Data used to generate this dataset include: (i) initial
void ratio, e0 = 0.955; (ii) swelling index with respect to
change in net normal stress, Cs = 0.088; (iii) swelling index
with respect to change in matric suction, Cm = 0.080; and
(iv) swelling pressure, Ps = 320 kPa. The two curves on the
extreme planes meet at a vertical net normal stress equal to
1 kPa and a suction equal to 1 kPa. The results are presented
in Fig. 9 using both a semilogarithmic scale and an arithmetic scale. This dataset corresponds to wetting unloading test
conditions.
SigmaPlot software (Systat 2000) was used for fitting the
equations to the data points. The R2 criterion and Akaike Information Criterion (AIC) (Akaike 1974) were used to evaluate the mathematical equations. The AIC criterion takes
into account the number of data points, the sum of the
weighted squared residuals, and the number of fitting param-

Soil properties characterization


Lytton (1994) presented typical values for coefficients of
earth pressure at-rest, K0, back calculated based on field observations of heave and shrinkage. A value of the coefficient
of earth pressure at rest equal to 0.67 was suggested for wetting conditions when cracks in the soil are essentially closed.
A Poissons ratio equal to 0.4 is calculated under this consideration from the following relation:
[22]

K0
1 + K0

Void ratio data associated with general three-dimensional


swelling were generated from the following parameters: initial void ratio, e0 = 0.955; swelling index with respect to
change in net normal stress, Cs = 0.088; swelling index with
respect to change in matric suction, Cm = 0.080; swelling
pressure, Ps = 320 kPa; and the two curves on extreme
planes meet at vertical net normal stress equal to 1 kPa and a
suction equal to 1 kPa.
Figure 12a presents a set of generated void ratios in a
semilogarithmic plot with an assumed Poissons ratio equal
to 0.4. The best-fit surface for the void ratio dataset is presented in an arithmetic plot in Fig. 12b. The degree of satu 2006 NRC Canada

Vu and Fredlund

1257

Fig. 8. Proposed procedure to solve the low net normal stress and suction problem.

Fig. 9. Generated data for the void ratio constitutive surface for Regina clay. e0 = 0.955; Cs = 0.088; Cm = 0.080; Ps = 320 kPa.

2006 NRC Canada

1258

Can. Geotech. J. Vol. 43, 2006

Table 1. Best-fit results of different void ratio equations to void ratio data for Regina clay.
ID

Function

Fitting parameters

Fredlund (1979)
Lloret and Alonso (1985)

e = a + b log ( ua ) + c log ( ua uw )
e = a + b log ( ua ) + c log ( ua uw )
+ d log ( ua ) log ( ua uw )
2
2



e = a + b ln 1 + + d ln 1 +
c
f

where a = a1 + a2(ua uw); b = b1 + b2 ln (ua uw);


d = d1 + d2 ln (ua uw)
e = a + b log [1 + ( ua ) + c( ua uw )]
1 + c ( ua ) + d ( ua uw )
e = a + b log

1 + f( ua ) + g ( ua uw )

a = 1.071; b = 0.027; c = 0.024


a = 1.186; b = 0.091; c = 0.083; d = 0.034

Fredlund (2000b)

Unsat-1
Unsat-6

Table 2. Statistical results of void ratio constitutive equations for


void ratio data for Regina clay.
ID

Number of fitting
parameters

R2

AIC

Fredlund (1979)
Lloret and Alonso (1985)
Fredlund (2000b)
Unsat-1
Unsat-6

3
4
8
3
6

0.5299
0.9469
0.9622
0.9890
0.9979

102
174
179
228
288

Note: AIC, Akaike Information Criterion.

ration surface can be calculated for general threedimensional swelling as shown in Fig. 12c.
The hydraulic conductivity function is predicted from the
saturated hydraulic conductivity and the SWCC (Fredlund
and Rahardjo 1993). Shuai (1996) presented the hydraulic
conductivity function for Regina clay using the Gardner
(1958) equation
[23]

kw =

kw0e

(u uw)
1+a a

wg

where kw0 = 0.4 108 m/s; e is void ratio; b = 18.5; a =


0.01; and n = 1.1.
The hydraulic conductivity constitutive surface is presented graphically in Fig. 12d for the void ratio surface as
shown in Fig. 12a.
The fitting results for the generated void ratio and degree
of saturation constitutive surfaces are shown in Table 3. The
coefficients of volume change associated with the generated
void ratio data (Fig. 12a) and an assumed degree of saturation surface (Fig. 12c) are presented in Fig 13. The elasticity
parameter functions are presented graphically in Fig 14.

Coupled and uncoupled analyses


Equations [1], [2], and [3] can be solved for three variables; namely, horizontal displacement, u; vertical displacement, v; and pore-water pressure, uw. Swelling is a timedependent process involving nonlinear soil properties. Initial
conditions must be specified for both coupled and uncoupled

a1 = 1.205; a2 = 5.2105; b1 = 0.724;


b2 = 0.292; c = 0.011; d1 = 0.705;
d2 = 0.287; f = 0.010

a = 1.186; b = 0.092; c = 0.610


a = 1.350; b = 0.095; c = 49.722;
d = 34.571; f = 4.38104;
g = 8.22104

analyses. Table 4 presents various parameters associated


with plane strain coupled and uncoupled approaches.
Coupled approach
In the coupled approach, the water phase continuity (i.e.,
seepage) equation and the equilibrium (i.e., stress deformation) equations are solved simultaneously; meaning that the
dynamic interdependence between the seepage and deformation problems is fully realized. Except for Poissons ratio, all
other soil parameters (i.e., E, H, Ew, Hw, and kw) are considered to be functions of both net normal stress and matric
suction. Boundary conditions for both the water continuity
equation (i.e., pore-water pressure and water flux) and equilibrium equations (i.e., displacements and load) must be defined. The results show displacements, induced stresses,
pore-water pressures, and water fluxes obtained at any time
during the transient process.
Uncoupled approach
In the uncoupled approach, the water phase continuity
(i.e., seepage) equation is solved independently from the
equilibrium (i.e., stress deformation) equations. The interdependent solutions of the equations is undertaken using an iterative manner where the flow portion of the formulation is
solved for a given time period and the resultant pore-water
pressure changes are used as input in a deformation analysis.
In turn, volume changes and induced stresses from the deformation analysis are used in the computation of the soil
properties for the next time period in the seepage analysis.
In the uncoupled approach, the dependent variables and
nonlinear soil properties are separated into two analyses;
namely, a seepage analysis and a stress-deformation analysis. For the seepage analysis, the dependent variable is porewater pressure. At each given time period, the elasticity parameters, E and H, for soil structure are calculated at initial
conditions of current period and assumed to remain unchanged over the time increment. Net normal stress is assumed to be unchanged in the seepage analysis; therefore,
the elasticity parameters, Ew and Hw, for the water phase, as
well as the hydraulic conductivity, kw, are taken to be functions of matric suction. Boundary conditions for the seepage
analysis can be either pore-water pressure (or hydraulic
head) or water flux. The results of the seepage analysis pro 2006 NRC Canada

Vu and Fredlund

1259

Fig. 10. Averaged AIC and R2 criterion for generated data for the void ratio constitutive surface.

Fig. 11. Best-fit void ratio constitutive surface using proposed unsat-6 function for generated void ratio data of Regina clay.

vide the pore-water pressure and water flux with time for the
time period under consideration.
The dependent stress-deformation variables are horizontal
displacement, u and vertical displacement, v. In addition to
Poissons ratio, two elasticity parameters, E and H, are required for the soil structure. These properties are a function
of matric suction when the initial net normal stress is unchanged. The elasticity parameters, Ew and Hw, for the water
phase and hydraulic conductivity, kw, are no longer needed
for stress deformation analysis. Boundary conditions for the
stress-deformation analyses can be of the displacement type
or load type. Results of the stress-deformation analysis provide displacements and induced stresses due to applied
boundary conditions and changes in pore-water pressure.
Solutions using the uncoupled approach depend on the
magnitude of selected time periods for the seepage analysis.
Short time periods allow the stress state in the soils and the
soil properties to be described more accurately. Consequently, the results provide more accurate values for porewater pressures and displacements.
Several approaches have been tested in the uncoupled
analysis. Differences arise from the stress state used for the
seepage and stress-deformation analyses. Uncoupled solutions obtained from four types of analyses are called UCS1,
UCS2, UCS3, and UCS4. The solution obtained using a cou-

pled approach is named as CS. The stress paths followed in


each type of analysis are shown in Figs. 15a and 15b for the
seepage analysis and the stress-deformation analysis, respectively. It should be noted that the initial stress state may vary
from one type of uncoupled analysis to another while boundary conditions remain the same.
In the UCS1 uncoupled analysis, the soil properties obtained for the extreme plane are used for the entire constitutive surface. Therefore, net normal stress is not taken into
consideration for both the seepage and stress-deformation
analysis. Volume changes and changes in stress were not
considered in the seepage analysis. A net mean stress of
0.78 kPa (i.e., the mean stress at which the volume change
index, Cm, was obtained) was used for the UCS1 uncoupled
analysis.
In the UCS2 uncoupled analysis, the variation in the stress
state in the soil profile under initial stress state conditions
was considered. Volume changes and changes in net normal
stress were not considered in the seepage analysis. Change
in net normal stress was not considered in the stressdeformation analysis.
In the UCS3 uncoupled analysis, the assumptions related
to the seepage analysis are the same as those used in the
UCS2 analysis; however, changes in net normal stress are
considered in the stress-deformation analysis.
2006 NRC Canada

1260

Can. Geotech. J. Vol. 43, 2006

Fig. 12. Constitutive surfaces for Regina clay.

In the UCS4 uncoupled analysis, changes in volume and


changes in net normal stress are considered in both the seepage and stress-deformation analysis. It should be noted that
variations in the stress state in the soil better reflect the
stress state conditions related to the problem.
The examples presented in this paper are analyzed using
various types of uncoupled analyses and the results are then
compared with the coupled solutions.

Computer programs
A finite element computer program, COUPSO, was used
to obtain the coupled solution and a general-purpose partial
differential equation solver, FlexPDE (PDE Solutions Inc.
2001), was used to obtain uncoupled solutions in this study.
The COUPSO program was developed by Pereira (1996) to
solve a coupled problem involving a small earth dam associated with unsaturated collapsing soils. Vu (2003) modified
the program to accommodate a coupled analysis for volume
change problems associated with an unsaturated expansive

soil. A detailed description of the COUPSO program and its


verification were presented by Pereira (1996) and Vu
(2003).
FlexPDE is a scripted finite element model builder and
numerical solver for both two- and three-dimensional problems. FlexPDE performs the operations necessary to turn a
description of the partial differential equation system into a
finite element model and then solves the system of equations. FlexPDE has automatic mesh generation and refinement, adaptive time step design, and time refinement. A
more detailed description of the FlexPDE program can be
found in PDE Solution Inc. (2001).

Computer results and discussions


Example 1: Infiltration of water from ground surface
This example considers the hypothetical case of a 5 m
thick deposit of swelling clay. The surface is partially covered with a flexible cover. Figure 16 presents the geometry
and key variable for this problem. The transient wetting pro 2006 NRC Canada

Vu and Fredlund

1261

Table 3. Fitting parameter results for general three-dimensional constitutive surfaces for Regina clay
Fitting parameters
Constitutive surface

Void ratio, e
Degree of saturation, S

1.2492
1.0000

0.0979
0.0725

4.8240
0.0125

3.3330
11.7265

0.0009
0.0125

0.0012
0.0071

Fig. 13. Coefficient of volume change functions for Regina clay.

cess is introduced by imposing a water infiltration rate equal


to 2 108 m/s at the uncovered portion of the ground surface. Such a wetting condition simulates water infiltration
into the soil mass due to the watering of a lawn or a light
rain. The analysis is performed to observe the swelling soil
behaviour and matric suction changes as the transient wetting front advances into the soil mass.
Figures 17 and 18 show the change of heave at ground
surface with time and the change of heave versus depth with
time, respectively, obtained from a coupled analysis. A max-

imum differential heave of about 30 mm can occur between


the cover location and the far point of the uncovered portion.
A maximum heave of about 58 mm could take place at the
surface and heave would essentially be complete after about
375 days of infiltration.
Figures 19 and 20 compare the changes of matric suction
and vertical displacement at the monitoring points in the soil
mass for various types of analysis. The matric suction below
the cover reduced rapidly in the first 50 days. The rate of
change in matric suction in the soil outside the cover re 2006 NRC Canada

1262

Can. Geotech. J. Vol. 43, 2006

Fig. 14. Elasticity parameter functions for Regina clay.

duced rapidly with time. Corresponding to the change in


suction, the horizontal stresses increased rapidly during the
first 50 days. Horizontal stresses increased more near ground
surface where the soil has a higher swelling potential. A total increase of about 58 kPa in horizontal stress was computed for point A. A total increase of about 20 kPa in the
horizontal stress was computed for point C.
The changes in matric suction at any elapsed time appear to
be overpredicted in the uncoupled analyses, resulting in the
differences in horizontal and vertical displacements. The uncoupled analyses in this study used elastic parameters at net
normal stresses that are lower than the actual net normal
stresses. The stiffness of an unsaturated expansive soil decreases with a decrease in net normal stress, resulting in a
larger amount of heave in uncoupled analyses. The magnitude
of the differences between uncoupled and coupled solutions,
therefore, depends on the stress paths followed in the uncou-

pled solutions. The uncoupled solution, UCS4 appear to be


essentially the same as that of the coupled solution, CS.
Figures 21, 22, and 23 present and compare the distributions of matric suction, horizontal displacement, and vertical
displacement, respectively, in the soil profile at day 53. Immediately after wetting was introduced below the uncovered surface, water flowed downward and to the left in the soil
domain. Matric suction near the uncovered surface reduced to
less than 100 kPa. Horizontal displacements decreased with
depth with a maximum value of 11 mm at ground surface
near to the cover. About 33 mm of heave took place at the uncovered location. The results of the analyses also show a significant increase in horizontal stress near ground surface,
particularly within the uncovered portion. An increase of
about 60 kPa in the horizontal stress was predicted for the soil
near ground surface where the soil had a high potential to
swell. There were increases in vertical stress because of the
2006 NRC Canada

Vu and Fredlund

1263

Table 4. Summary of uncoupled and coupled model for two-dimensional swelling analysis associated with an expansive soil.
Uncoupled approach
Description

Seepage model

Stressstrain model

Coupled approach

Governing PDE(s)

Water continuity equation

Stress equilibrium equations

Computer program
Dependent variables

FlexPDE
Pore-water pressure, uw

FlexPDE
Horizontal displacement, u
Vertical displacement, v

Initial conditions

(mean ua)i
(ua uw)i

E, H at [(mean ua)i ; (ua uw)i]


Ew = fn(ua uw) at (mean ua)i
Hw = fn(ua uw) at (mean ua)i
kw = kw(ua uw) at (mean ua)i

(mean ua)i
(ua uw)i

Stress equilibrium equations


Water continuity equation
COUPSO
Horizontal displacement, u
Vertical displacement, v
Pore-water pressure, uw
(mean ua)i
(ua uw)i

Boundary conditions (valued)

Pore-water pressure, uw

Displacement, u, v

Boundary conditions (natural)

Water flux, q

Applied load

Output

Pore-water pressure, uw

Displacements, u, v
Resulting stresses

Analysis for

A change in time, t

A change in suction, uw

Soil properties as constants


Soil properties as functions

E = fn(mean ua) at (ua uw)i


H = fn(mean ua) at (ua uw)i

E = fn[(mean ua), (ua uw)]


H = fn[(mean ua), (ua uw)]
Ew = fn[(mean ua), (ua uw)]
Hw = fn[(mean ua), (ua uw)]
kw = fn[(mean ua), (ua uw)]
Displacement, u, v
Pore-water pressure
Applied load
Water flux
Displacements, u, v
Pore-water pressure, uw
Resulting stresses
Volumetric water content
A change in time, t

Fig. 15. Stress path followed in seepage model and stress-deformation model of various types of uncoupled analyses.

Fig. 16. Illustration of the geometry and key variables for Example 1: Infiltration of water from ground surface.

2006 NRC Canada

1264

Can. Geotech. J. Vol. 43, 2006

Fig. 17. Change of heave at ground surface with time, Example 1: Infiltration of water from ground surface, coupled solution.

Fig. 18. Change of heave versus depth with time, Example 1: Infiltration of water from ground surface, coupled solution.

Fig. 19. Comparison of matric suction development with time for points A, B, and C; Example 1: Infiltration of water from ground
surface.

2006 NRC Canada

Vu and Fredlund

1265

Fig. 20. Comparison of vertical displacement development with time for points A, B, and C; Example 1: Infiltration of water from
ground surface.

Fig. 21. Comparison of matric suction distribution at day 53; Example 1: Infiltration of water from ground surface.

Fig. 22. Comparison of horizontal displacement distribution at day 53; Example 1: Infiltration of water from ground surface.

increase of the soil self weight due to saturation, but these


changes were insignificant in comparison with the change in
horizontal stress. Most of the increases in void ratio and degree of saturation took place within the uncovered portion
near to the ground surface where infiltration took place.

Example 2: Leakage of water below a flexible cover


Example 2 considers the hypothetical case of a 5 m layer
of swelling clay below a flexible cover (Fig. 24). Initial
matric suction is taken to be constant throughout the depth
and equal to 400 kPa. It is assumed that a leaking water line
2006 NRC Canada

1266

Can. Geotech. J. Vol. 43, 2006

Fig. 23. Comparison of vertical displacement distribution at day 53; Example 1: Infiltration of water from ground surface.

Fig. 24. Illustration of the geometry and key variables for Example 2: Leakage of water below a flexible cover.

Fig. 25. Change of heave at ground surface with time, Example 2: Leakage under a flexible cover, coupled solution.

produced zero pore-water pressure under the cover. Deformation and matric suction profiles versus time were computed.
Figures 25 and 26 show the change of heave at ground
surface with time and the change of heave versus depth with
time, respectively, obtained from a coupled analysis. The
heave patterns indicate that most of the heave below the

cover occurred in the first 100 days after wetting commenced. Most of the differential heave (i.e., about 65 mm)
took place in the first 56 days. In the period from day 56 to
day 150, heave increased gradually at the same rate throughout the entire soil mass. After this period, heave developed
faster along the right side of the soil mass. A maximum
heave of 112 mm was predicted at day 350.
2006 NRC Canada

Vu and Fredlund

1267

Fig. 26. Change of heave below the cover versus depth with time, Example 2: Leakage under a flexible cover, coupled solution.

Fig. 27. Comparison of matric suction development with time; Example 2: Leakage under a flexible cover.

Fig. 28. Comparison of horizontal displacement development with time; Example 2: Leakage under a flexible cover.

2006 NRC Canada

1268

Can. Geotech. J. Vol. 43, 2006

Fig. 29. Comparison of vertical displacement development with time; Example 2: Leakage under a flexible cover.

Fig. 30. Comparison of matric suction distribution at day 150; Example 2: Leakage under a flexible cover.

Fig. 31. Comparison of horizontal displacement distribution at day 150; Example 2: Leakage under a flexible cover.

Figures 27, 28 and 29 compare the changes of matric suction, horizontal displacement, and vertical displacement, respectively, at a selected point in the soil mass for various
types of analyses. Figure 27 shows that water appeared to
move slightly faster in the uncoupled analyses, when

changes in stress and deformation were not considered in the


seepage equation (i.e., UCS2 and UCS3). The changes in
matric suction at any elapsed time were overpredicted in
these analyses resulting in slight differences in horizontal
and vertical displacements. It should be noted that changes
2006 NRC Canada

Vu and Fredlund

1269

Fig. 32. Comparison of vertical displacement distribution at day 150; Example 2: Leakage under a flexible cover.

in stress were not considered in solution UCS2, but were


considered in solutions UCS3 and UCS4. The UCS4 uncoupled solution compares well with the coupled solution, CS.
Figures 30, 31, and 32 compare the distribution of matric
suction, horizontal displacement, and vertical displacement,
respectively, in the soil profile at day 150. At day 150, the
matric suction value below the cover reduced to less than
50 kPa. The wetting front reached the extreme ends of the
soil mass. Matric suctions varied from 0 kPa to 300 kPa in
the soil mass at this time. Cumulative horizontal and vertical
displacements had the same patterns as those at day 56;
however, soil in the right portion of the soil mass was
pushed to the left as matric suctions reduced significantly in
this portion. A maximum cumulative heave of 105 mm took
place at the centre of the cover. About 25 mm of heave could
be observed at the far side of the soil mass.

Concluding remarks
The difficulties associated with numerical modelling of an
unsaturated expansive soil have been presented and discussed. The characterization of the swelling portion of the
void ratio surface for an unsaturated expansive soil from
swelling indices provides a practical approach for engineering practice. Two mathematical equations were proposed to
describe the void ratio constitutive surface even when the net
normal stress and matric suction approached zero values.
The equations were shown to successfully characterize the
constitutive surfaces for both the soil structure and water
phase of Regina clay.
Solutions to the two example problems associated with
unsaturated expansive soils have been obtained using both
the uncoupled and the coupled analyses. Several types of uncoupled analyses, arising from various assumptions related
to the stress path followed in the water flow analysis and
stress-deformation analysis in the uncoupled solutions have
been presented. The coupled analysis generally produces
slightly smaller displacements when compared to the uncoupled analyses. Changes in the stress and deformation associated with the water flow analysis in the swelling soils should
be taken into account as the soil approaches saturation. The
effect of the changes in stress and deformation appeared to
be minor when the soil has high matric suctions. The coupled seepage and stress-deformation analysis provides a

more rigorous understanding of the swelling behaviour of


expansive soils and forms a reference for the evaluation of
various uncoupled analyses. When changes in net normal
stress and volumetric deformation are considered in an uncoupled analysis, the uncoupled and coupled solutions provide similar results.

References
Akaike, H. 1974. A new look at the statistical model identification.
Institute of Electrical and Electronics Engineers Transactions on
Automatic Control, 19(6): 716723.
Alonso, E.E, Gens, A., and Josa, A. 1990. A constitutive model for
partially saturated soils. Gotechnique, 40(3): 405430.
Al-Shamrani, M.A., and Al-Mhaidib, A.I. 2000. Swelling behaviour under eodometric and triaxial loading conditions. In Advances in Unsaturated Geotechnics: Proceedings of Sessions of
Geo, Denver, Colo., 58 August 2000. Edited by C.D.
Shackelford, S.L. Houston, and N.-Y. Chang. American Society
of Civil Engineers, Geotechnical Special Publication 99, Reston,
Va. pp. 344360.
Biot, M.A. 1941. General theory of three-dimensional consolidation. Journal of Applied Physics, 12(2): 155164.
Casagrande, A. 1936. The determination of the preconsolidation
load and its practical significance. Discussion D-34. In Proceedings of the 1st International Conference on Soil Mechanics and
Foundation Engineering, Cambridge, Mass. Vol. 3, pp. 6064.
Escario, V. 1969. Determination of the geotechnical characteristics
of expansive soils. In Proceedings of the 2nd International Conference on Expansive Soils, Texas A&M University, 1719 August 1969. Edited by S.J. Buchanan. Texas A&M Press, College
Station, Tex. pp. 114120.
Fredlund, D.G. 1979. Second Canadian Geotechnical Colloquium:
Appropriate concepts and technology for unsaturated soils. Canadian Geotechnical Journal, 16(1): 121139.
Fredlund, D.G. 2000a. The 1999 R.M. Hardy Lecture: The implementation of unsaturated soil mechanics into geotechnical engineering. Canadian Geotechnical Journal, 37(5): 963986.
Fredlund, D.G., and Rahardjo, H. 1993. Soil mechanics for unsaturated soil. John Wiley & Sons, New York, 560 p.
Fredlund, D.G., and Xing, A. 1994. Equations for the soil-water
characteristic curve. Canadian Geotechnical Journal, 31(4):
521532.
Fredlund, M.D. 2000b. The role of unsaturated soil property functions in the practice of unsaturated soil mechanics. Ph.D. dissertation, University of Saskatchewan, Saskatoon, Sask.
2006 NRC Canada

1270
Gardner, W.R. 1958. Some steady state solutions of the unsaturated
moisture flow equation with application to evaporation from a
water table. Soil Science, 85: 228232.
Gens, A., and Alonso, E.E. 1992. A framework for the behaviour
of unsaturated expansive clays. Canadian Geotechnical Journal,
29:10131032.
Graham, J., and Li, E.C.C. 1985. Comparison of natural and
remolded plastic clay. Journal of Geotechnical Engineering,
American Society of Civil Engineers, 111(7): 865881.
Ho, D.Y.F., Fredlund, D.G., and Rahardjo, H. 1992. Volume
change indices during loading and unloading of an unsaturated
soil. Canadian Geotechnical Journal, 29(2): 195207.
Holtz, W.G., and Gibbs, H.J. 1956. Engineering properties of expansive clay. Transactions of the American Society of Civil Engineers, 121: 641663.
Lloret, A., and Alonso, E.E. 1985. State surfaces for partially saturated soils. In Proceedings of the 11th International Conference
on Soil Mechanics and Foundation Engineering, San Francisco,
Calif., 1216 August 1985. Edited by Publication Committee of
XI ICSMFE, A.A. Balkema, Rotterdam, The Netherlands.
Vol. 2, pp. 557562.
Lytton, R.L. 1994. Prediction of movement in expansive clay. In
Vertical and Horizontal Deformations of Foundations and Embankments: Proceedings of Settlement 94, College Station,
Texas, 1618 June 1994. Edited by A.T. Yeung and G.Y. Feaalio.
American Society of Civil Engineers, Geotechnical Special Publication 40, pp. 18271845.
Matyas, E.L. 1969. Some properties of two expansive clays from
western Canada. In Proceedings of the 2nd International Conference on Expansive Soils. Texas A&M University, College Station, Texas, 1719 August 1969. Edited by S.J. Buchanan. Texas
A&M Press, College Station, Tex. pp. 263278.
Matyas, E.L., and Radhakrishna, H.S. 1968. Volume change characteristics of partially saturated soils. Gotechnique, 18(4): 432
448.
PDE Solutions Inc. 2001. FlexPDE 3 Reference manual. PDE Solutions Inc., Antioch, Calif.
Pereira, J.H.F. 1996. Numerical analysis of the mechanical behavior of collapsing earth dams during first reservoir filling. Ph.D.
dissertation, University of Saskatchewan, Saskatoon, Sask.
Pham, Q.H., Fredlund, D.G., and Padilla, J.M. 2004. Use of the
GCTS apparatus for the measurement of soil-water characteristic curve. In Proceedings of the 57th Canadian Geotechnical
Conference, Quebec City, Que. 2427 October 2004. pp. 16.
Richards, B.G., Peter, P., and Martin, R. 1984. The determination
of volume change properties in expansive soils. In Proceedings
of the 5th International Conference on Expansive Soils,
Adelaide, Australia, 2123 May 1984. Edited by P.W. Mitchell,
Organized by the Australian Geomechanics Society and the Institution of Engineers, Australia, National Conference Publication No. 84/3. pp. 179186.
Shuai, F. 1996. Simulation of swelling pressure measurements on
expansive soils. Ph.D. dissertation, University of Saskatchewan,
Saskatoon, Sask.
Systat Software Inc. 2000. Sigmaplot users guide. Systat Software
Inc., Point Richmond, Calif.
Terzaghi, K. 1943. Theoretical soil mechanics. John Wiley & Sons,
New York.
Vu, H.Q. 2003. Uncoupled and coupled solutions of volume change
problems in expansive soils. Ph.D. thesis, Department of Civil
Engineering, University of Saskatchewan, Saskatoon, Sask.
Vu, H.Q., and Fredlund, D.G. 2004. The prediction of one-, two-,
and three-dimensional heave in expansive soils. Canadian
Geotechnical Journal, 41: 713737.

Can. Geotech. J. Vol. 43, 2006

List of symbols
a, b, c, d, e, f, g fitting parameters
bx, by body forces in x- and y-directions, respectively,
[ML2T2]
Cm swelling index, slope of the void ratio versus
logarithm of matric suction
Cs swelling index, slope of the void ratio versus
logarithm of net normal stress
e void ratio of the soil, [decimal]
ef initial void ratio of the soil at a net normal stress
equal to ( ua)0 and a matric suction equal to
s0, [decimal]
e0 initial void ratio of the soil, [decimal]
E elasticity parameter function for the soil structure with respect to a change in net normal
stress, [ML1T2]
Ew elasticity parameter function for the change in
the amount of water in the soil with respect to a
change in net normal stress, [ML1T2]
g gravitational acceleration, [LT2]
Gs specific gravity
H elasticity parameter function for the soil structure with respect to a change in matric suction,
[ML1T2]
Hw elasticity parameter function for the change in
the amount of water in the soil with respect to a
change in matric suction, [ML1T2]
x
y
,
kw kw the hydraulic conductivities in the x- and y-directions, respectively, [LT1]
K0 coefficient of earth pressure at rest
m1w coefficient of water volume change with respect
to a change in net normal stress, [M1LT2]
m2w coefficient of water volume change with respect
to a change in matric suction, [M1LT2]
s
m1 coefficient of volume change with respect to a
change in net normal stress, [M1LT2]
m2s coefficient of volume change with respect to a
change in matric suction, [M1LT2]
Ps swelling pressure, [ML1T2]
s0 matric suction at which the swelling index Cs
was obtained, [ML1T2]
ss swelling suction (i.e., swelling pressure equivalent value of on the suction plane, [ML1T2]
S degree of saturation, [decimal]
Si initial degree of saturation, [decimal]
t time, [T]
u, v displacements in the x- and y-directions, respectively, [L]
ua pore-air pressure, [ML1T2]
uw pore-water pressure, [ML1T2]
(ua uw) matric suction, [ML1T2]
(ua uw)i initial matric suction, [ML1T2]
V0 initial overall volume, [L3]
Vw volume of water, [L3]
x, y, z rectangular Cartesian coordinates
Y elevation, [L]
t total or moist unit weight, [ML2T2]
v volumetric strain
w volumetric water content, [decimal]
Poissons ratio
2006 NRC Canada

Vu and Fredlund

1271

w density of water, [ML3]


x , y , z net total stresses in x-, y-, and z- direction, respectively, [ML1T2]
mean mean net total stress, mean = ( x + y + z)/3,
[ML1T2]
( ua) net normal stress, [ML1T2]
( ua)i initial net normal stress, [ML1T2]
( ua)0 net normal stress at which the swelling index
Cm was obtained, [ML1T2]

The swelling suction, ss, can be estimated from eq. [A3]


as follows:
Cs / Cm

[A4]

Ps

ss = s0

( ua ) 0

Let us assume that e is the void ratio at a net normal stress


equal to X, and matric suction equal to Y. The change in void
ratio (Fig. A1) can be written as
[A5]

e = e e0

Appendix A. Derivation of equation for the


estimation of void ratio surface from
volume change indices

From the net normal stress plane at a matric suction equal


to b, (Fig. 5), the change in void ratio can be expressed as
follows:

The void ratio, ef, at a net total stress equal to ( ua)0,


and a matric suction equal to s0, can be calculated from either the net normal stress plane or the matric suction plane,
as shown in Fig. A1 and the following equations.

[A6]

[A1]

( ua ) 0
ef = e0 + C s log
Ps

[A2]

ef = e0 + C m log

s0
ss

where Ps is swelling pressure; ss is swelling suction (i.e.,


equivalent value of swelling pressure on the suction plane).
Suction s0 and net total stress ( ua)0 are values of
stress points at which the void ratio versus logarithm of suction and the void ratio versus logarithm of net normal stress
converge (Fig. A1). These values are measurable in a laboratory program. A value of 1 kPa was assumed for the reference stress states of suction s0 and net total stress (
ua)0 in this study.
Equating eqs. [A1] and [A2] gives
[A3]

C s log

( ua ) 0
s
= C m log 0
Ps
ss

e = C s log

M
Ps

Rearranging eq. [A6], gives


[A7]

M = Ps 10e / Cs

where ( ua)0 < M < Ps

From the matric suction plane at a net normal stress equal


to ( ua)0 (Fig. A1), a change in void ratio can be written
as follows:
[A8]

e = C m log

N
ss

Rearranging eq. [A8] gives


[A9]

N = ss 10e / Cm

where s0 < N < Ss

For a constant void ratio plane at a void ratio equal to e,


the following relationship can be obtained (Fig. A2):
[A10]

N s0
Y s0
=
M ( ua ) 0 M X

Rearranging eq. [A10] gives


[A11] Y =

N s0
( M X ) + s0
M ( ua ) 0

Substituting eqs. [A4], [A5], [A7], and [A9] into


eq. [A11] gives the relationship amongst net normal stress X,
matric suction Y, and void ratio, e
P Cs / Cm

s0 s
10 ( e e0 )/ Cm 1 (Ps 10 ( e e0 )/ Cs X )
a

[A12] Y =
+ s0
( e e0 )/ Cs
Ps 10
( ua ) 0
Equation [A12] can be rewritten setting the X variable to net normal stress and the Y variable to matric suction
P Cs / Cm

s0 s
10 ( e e0 )/ Cm 1 [Ps 10 ( e e0 )/ Cs ( ua )]
a

[A13] (ua uw) =


+ s0
( e e0 )/ Cs
Ps 10
( ua ) 0

2006 NRC Canada

1272

Can. Geotech. J. Vol. 43, 2006

Fig. A1. Linear relationships between void ratio and logarithm of stress state variables at extreme planes.

Fig. A2. Relationship between net normal stress and matric suction at constant void ratio plane.

2006 NRC Canada

You might also like