You are on page 1of 13

Proc. of Inter.

Seminar on Practice and Advance in Geotechnical Engineering, Shanghai, China

BEHAVIOR OF REINFORCED EMBANKMENT OVER SOFT SUBSOIL

J.-C. CHAI
Institute of Lowland Technology, Saga University, Saga, Japan
H.-H. Zhu
Department of Geotechnical Engineering, Tongji University, Shanghai, China

ABSTRACT: The functions of reinforcement on the behavior of reinforced embankment over


soft subsoil are discussed first. The mobilized tensile force in the reinforcement contributes to
the stability of the embankment, and the confining effect of the reinforcement to embankment
fill material and soft subsoil can increase the bearing capacity and reduce the lateral
deformation of soft subsoil. However, the effect of the reinforcement on soft subsoil
deformation can only become significant when soft subsoil approaches to failure. Then some
techniques for modeling the reinforced embankment on soft subsoil are presented, namely,
modeling the soil/reinforcement interaction behavior, embankment construction process and
hydraulic conductivity variation during consolidation process of soft subsoil. It is
demonstrated that these factors are important on simulating the behavior of the reinforced
embankment on soft subsoil. Finally, two case histories of the reinforced embankment on soft
subsoil are presented. One of the embankments is a geogrid reinforced embankment on Muar
clay deposit in Malaysian and another one is a built-to-failure geotextile reinforced
embankment in Lian-Yun-Gang, China.

INTRODUCTION

In case of constructing an embankment over soft subsoil of low strength and high
compressibility, the engineering tasks are to prevent the failure of the embankment and to
control the subsoil deformation within an allowable limit. Several methods have been
developed for economically and safely constructing embankment on soft subsoil. Using basal
reinforcements is one of the methods.
The mechanisms of reinforcing an embankment over soft subsoil are discussed elsewhere
(Bonaparte and Christopher 1987; Jewell 1988). The functions of basal reinforcement are: (1)
increasing the stability of embankment by mobilized tensile force in the reinforcement, and
(2) providing a confinement to embankment fill and foundation soil adjacent to the
reinforcement. This confining effect can reduce the lateral distortion to subsoil due to
embankment load, and therefore, the shear stress in soft subsoil. However, the mobilization of
these functions depends on the soil/reinforcement interaction behavior, which relates to the
strength of reinforcement, the relative stiffness of reinforcement and subsoil, embankment
geometry, and the factor of safety (FS) of the embankment at working state.
This paper first describes the functions of basal reinforcement for an embankment over
soft subsoil. Then some techniques of modeling the reinforced embankment on soft subsoil
are discussed. Finally, two case histories of reinforced embankment on soft subsoil are
presented to illustrate the magnitude of the reinforcement effects on FS of the embankment
and subsoil deformation.

- 71 -
FUNCTIONS OF REINFORCEMENT

Effect of reinforcement on FS of embankment

Practically, the limiting equilibrium method is used in analyzing FS of a reinforced


embankment on soft subsoil. In most proposed methods (e.g. Milligan and Rochelle 1984;
Mylleville and Rowe 1988), it is assumed that the reinforcement does not change the failure
surface much and only considers the restoring moment due to the mobilized reinforcement
tensile force. The FS of a reinforced embankment can be expressed as the follows.
Re storing Moments MRR + MRSOIL
FS = = (1)
Overturing Moments MOSOIL
where MRR is restoring moment due to tensile force developed in the reinforcement, MRSOIL
is restoring moment due to the mobilized shear strength of soil along the failure surface, and
MOSOIL is overturning moment due to embankment fill. The effect of MRR on FS depends
on the magnitude of MRR and the ratio of MRR/MRSOIL. For a given value of MRR, the
larger the embankment (larger the MRSOIL), the smaller the MRR/MRSOIL, and the less
effect of MRR on FS.
The magnitude of MRR depends on the mobilized tensile stress in reinforcement (T). T
depends on both mobilized tensile strain in reinforcement and the reinforcement stiffness.
Rowe and Soderman (1985) discussed the strain compatibility between reinforcement and soil
and pointed out that the allowable compatible strain, which is the tensile strain for design
reinforcement, is a function of both subsoil properties and the geometry of embankment.
From their finite element analysis results, the allowable compatible strain is in a range of 1%
to 9%. By reviewing some of the published data, Bonaparte and Christopher (1987) suggested
that the allowable strain is mainly controlled by subsoil properties. The recommended values
are 2~3% for sensitive subsoil, 4~6% for medium to low sensitive subsoil, and 6~10% for
non-sensitive plastic subsoil.
For geogrids and woven geotextiles, the confined in-soil stiffness is not different much
from the in-air value. However, for non-woven geotextiles, it is generally agreed that the soil
confinement increases the friction resistance between fabrics of geotextile and therefore, the
stiffness of a geotextile. Miura and Chai (1999) proposed a method for determining the in-soil
stiffness of a geotextile by combining a small-scale soil/geotextile interface shear test and a
large-scale pullout test results. For the geotextile tested, for tensile strain less than 5%, the
in-soil stiffness was 2~3 times of in-air one. Beyond this limiting strain, the in-soil tangent
stiffness was the same as the in-air value.
There has been debate on the orientation of the mobilized tensile force in the reinforcement,
i.e. horizontal or tangent to the failure surface. Rowe and Li (2002) stated that there is strong
evidence that the reinforcement tensile force should be taken to act in its original horizontal
orientation.
In calculating the FS of a reinforced embankment on soft subsoil, evaluating the
mobilized undrained shear strength (Su) of soft subsoil is also an important issue. Ladd (1991)
proposed an empirical equation for estimating Su of soft clay as follows:

S u = S (OCR ) m v (2)

where S and m are constants, OCR is overconsolidation ratio, and v is vertical effective
stress. Normally, the value of S is in a range of 0.2~0.4, and m is from 0.75 to 1.0. If S value
can be determined by local experience or back calculated from some case histories, Eq. 2
provides a reasonable estimation of Su of soft subsoil.
For embankment construction, the critical time is the end of construction. During
construction, the partial drainage effect can increase the effective stress in subsoil and
therefore the value of Su, especially for prefabricated vertical drain (PVD) improved subsoil.

- 72 -
Chai and Miura (2000) developed a CL CL
one-dimensional finite element
program to predict the value of Su of Thrust
Thrust
pressure
PVD improved subsoil at the end of pressure
Embankment Embankment
embankment construction.
Soft subsoil
Effect of reinforcement on subsoil Soft subsoil
response Reinforcement

(a) Without reinforcement (b) With reinforcement


As shown in Fig. 1(a) that in case Fig. 1 Illustration of the effect of reinforcement
of without reinforcement, the lateral on subsoil
spreading of fill material will induce
the outward shear stress () on the
surface of the soft subsoil, which
will cause the reduction of the
bearing capacity and increase the
lateral displacement of the soft
subsoil. In case of reinforced
embankment, the reinforcement
provides a confinement to the
embankment fill and the soft subsoil
(Fig. 1(b)) and reduces the lateral
displacement of the soft subsoil.
Fig. 2 Nc values for nonhomogenous soil (after
Bearing capacity Rowe and Soderman 1987)
With a high stiffness and strength
reinforcement, the slip surface is
difficult to pass through the
embankment itself and the failure
mechanism of the embankment will
be the bearing capacity failure of soft
subsoil. The ideal (or ultimate)
condition is that the reinforced
embankment behaves like a rigid
footing. From general bearing Fig. 3 Bearing capacity calculation of a perfectly
capacity theory, the ultimate bearing reinforced embankment (after Rowe and Soderman
capacity (qu) of a rigid footing on 1987)
soft subsoil can be expressed as:
qu = N c S u 0 + q s (3)
where Nc is bearing capacity factor, Su0 is initial undrained shear strength of soft subsoil, and
qs is distributed surcharge load on the surface of subsoil. The strength of natural clay deposit
is not uniform and normally increases with depth. Davis and Booker (1975) published a
solution considering the effect of strength increase with depth. For rough footing case, Rowe
and Soderman (1987) provided a graphic form (Fig. 2) for obtaining the bearing capacity
factor, Nc, based on Davis and Bookers solution.
Under embankment loading, the loading intensity varies under the shoulder. Rowe and
Soderman (1987) proposed an approximate method to convert the embankment load to an
equivalent rigid footing as illustrated in Fig. 3. From plasticity solution, the pressure at the
edge of a rigid footing is (2+)Su0. Under embankment shoulder, there is a point with a fill
thickness of h*, at which the following condition holds:
h* = (2 + ) S u 0 (4)
where is unit weight of embankment fill. Then the effective width (b) of the equivalent

- 73 -
footing will be between these kind points on
either side of the embankment and can be
expressed as the following:
b = B + n( H h*) (5)

where B is top width and n is cotangent of the


embankment slope angle. The equivalent
loading will be the average embankment fill
loading within b. The embankment loads
outside the effective width (b) are considered as
surcharge loads and averaged within a distance
x from the edge of the footing. The value of x
Fig. 4 Determine the failure zone
approximately equals to the depth (d) to which
beneath a rough rigid footing (after
the failure mechanism is expected to extend and
Rowe and Li 2002)
can be estimated from Fig. 4 (after Rowe and Li
2002). At the either side of the embankment toe,
the embankment width will be considered as P-2
surcharge loading is nh*. In case that nh* > x, P-1 2
2

Devistor Stress
only the loads within x will be considered. 1
1

Subsoil deformation.
The confinement from the reinforcement to 1= 2 =
soft subsoil can reduce the lateral spreading of 1<< 2
the subsoil and which can contribute to reduce
the settlement. Rowe and Li (2002) stated that
for a particular geometry of embankment and
Shear Strain
soil profile, there is a threshold reinforcement
stiffness below which the reinforcement has no Fig. 5 Illustration of the non-linear
effect on settlement. However, Chai et al. shear stress/strain response (after Chai
(2002) demonstrated that stress level in subsoil et al. 2002)
is a most crucial factor controlling the effect of
the reinforcement on subsoil deformation. It is
concluded that only when the embankment load reaches the level at which the unreinforced
case very close to failure (FS=1.0), the reinforcement will have obvious effect on subsoil
deformation. At working condition of FS=1.2 to 1.3, the effect of reinforcement on subsoil
deformation is small and insignificant. Figure 5 is used to illustrate why the effect of
reinforcement on subsoil deformation increases when embankment approaches failure.
Normally, the stress-strain curve of soft clay soils is non-linear. Assuming that the
reinforcement can reduce certain amount of shear stress () on a soil element. At P-1, with a
higher FS (about 1.2), the reduction of shear stress can result in a reduction of shear strain
of 1. However, at P-2 with a value of FS close to unity, can result in a larger shear strain
reduction 2 (2>>1).

MODELING REINFORCED EMBANKMENT ON SOFT SUBSOIL

The behavior of the reinforced embankment on soft subsoil depends on soil/reinforcement


interaction as well as fill/soft subsoil interaction. Finite element method (FEM) is suitable for
analyzing this kind of problem. However, the accuracy of the finite element analysis not only
depends on the constitutive models and the parameters used but also on the numerical
techniques adopted, such as selecting the proper soil/reinforcement interface properties, the
methods of applying the embankment load and simulating the hydraulic conductivity
variation during consolidation process of soft subsoil.

- 74 -
Modeling soil/reinforcement interaction

Soil/reinforcement interface property is


one of the important parameters that
influence the behavior of the embankment.
Soil/reinforcement interaction mode can
either be direct shear or pullout (Fig. 6, after
Hird and Kwok 1989). For grid or strip
reinforcements, these two different
interaction modes will yield different
interface strength and deformation
parameters. Usually, the direct-shear mode
will yield higher interface strength than the Fig. 6 Soil/reinforcement interaction
pullout mode (My11evi11e and Rowe 1988; modes (after Hird and Kwok 1989)
Chai and Bergado 1993b). The finite
element technique should be able to automatically select proper interface properties according
to the interaction modes. The technique proposed by Chai and Bergado (1993b) considers the
interface elements above and below the reinforcement as pair elements, and the signs of the
shear stresses of the pair elements are compared to determine whether the direct shear (same
sign) or the pullout (different sign) is the acting mode.

Modeling the construction process

The actual embankment construction is carried out by placing and compacting fill material
layer by layer. In finite element analysis, the mesh for embankment is pre-defined and the
embankment load is applied by turning on the gravity force of the embankment elements layer
by layer. If the construction settlement is not significant or if the embankment height (not fill
thickness) at the end of construction is specified, this process is satisfactory. However, for
embankments on soft subsoil, if the fill thickness is specified during the analysis, such as
predicting the embankment behavior under a given fill thickness, this process will yield a
wrong result. The analysis of embankment on soft subsoil is a large deformation problem
because amount of settlement at the end of construction can vary from 20 to 100% of the total
settlements (Asaoka et a1. 1992). The large deformation phenomenon can be approximately
considered by updating the nodal coordinates during the incremental analysis. Usually, this
operation does not include the elements above the current construction level. As a result, the
finally applied embankment fill thickness will be larger than pre-defined value. There are two
methods to treat this problem.
(1) Include all the embankment elements into all steps of incremental analysis and apply
the gravity force by percentage in each step
(Britto and Gunn 1987).
(2) Apply the embankment elements layer
by layer and correct the coordinate of the
elements above the current construction level
according to the settlement of the current
construction top surface as illustrated in Fig. 7
(after Chai and Bergado 1993a).
In case of an elastic ground, the above
two methods may not yield much different
results, but soft subsoil behaves
elasto-plastically and these two methods give
a quite different results. For a reinforced Fig. 7 Configuration of an embankment
embankment on soft Muar clay deposit, during construction (after Chai and
Malaysia, the settlement profiles obtained by Bergado 1993a)

- 75 -
the above two approaches are compared in
Fig. 8 (after Chai and Bergado 1993b). It
can be seen that the method of applying the
embankment elements layer by layer and
correct the coordinate of un-constructed
elements yielded a much better simulation
of the field behavior. The details of the
embankment will be described in next
section.

Modeling hydraulic conductivity


variation
Fig. 8 Comparing the settlement profiles
Hydraulic conductivity (k) is one of the (after Chai and Bergado 1993b)
main factors controlling the consolidation
process of soft subsoil. The value of k
depends on the size, shape and distribution 10-5

Coefficient of consolidation
of voids of soil. Taylor (1948) proposed a Ariake clay
relation between k and the void ratio (e) of
clay as follows. C (m /sec) 10-6
2

k = k 0 10 ( e0 e ) / Ck (6)
10-7
where k0 is initial hydraulic conductivity, e0 k varies (e-ln(p'))
k varies (ln(e)-ln(p'))
is initial void ratio, k is current hydraulic k constant (e-ln(p'))
conductivity, e is current void ratio, and Ck 10-8 0 Measured 1
10 10 102 103
is a constant (Ck=0.5e0, Tavenas et al. Vertical effetive stress, kPa
1983).
The coefficient of consolidation (C) is a Fig. 9 Variation of coefficient of
function of k and the coefficient of volume consolidation (modified from Chai 2001)
change (mv). Most elasto-plastic soil models
for clay use a linear e-ln(p) relation (p is effective mean stress). For structured natural clay,
Chai (2001) proposed a linear ln(e+ec)-ln(p) relation (ec is a constant). From these
relationships, the value of mv can be evaluated. Fig. 9 shows an example comparison of the

Table 1. Cf Value for a few clay deposits


Method for evaluation
Deposit Cf value field of hydraulic Remarks
conductivity
Bangkok clay at Asian Institute of 25 Back-analysis Chai et al. (1995)
Technology campus (about 100 km
from sea)
Bangkok clay at Nong Hao (close to 4 Back-analysis Chai et al. (1996)
sea)
Malaysia Muar clay deposit 2 Back-analysis Chai and Bergado
(1993a)
Ariake clay (close to sea area) 4 Back-analysis Chai and Miura (1999)
Louiseville (Canada) About 1a Self-boring permeameter Tavenas et al. (1986)
St-Alban (Canada) About 3a Self-boring permeameter Tavenas et al. (1986)
6
Soft mucky clay (eastern China) Back-analysis Shen et al. (2000)
a
Laboratory value was determined by direct measurement. For other cases, laboratory values were deduced
from Cv value (Cv is the vertical coefficient of consolidation).

- 76 -
variation of C with consolidation pressure under several different assumptions. In the figure,
k varies means that k varies with void ratio according to Eq. 6. It is very clear that if not
considering the k variation with void ratio, C will be over estimated.
Another factor is that laboratory test normally under-estimates the field k value because of
sample disturbance and sample size effect. If define the ratio between the field value and the
laboratory value as Cf ( C f = k field / k laboratory ), some available data of Cf are summarized in Table
1. It is recommended that the field value of k should be obtained by back-analysis of case
histories or the field hydraulic
conductivity measurement.

TWO CASE HISTORIES

A geogrid reinforced embankment


on Muar clay deposit, Malaysia, and
built to failure embankments (with and
without reinforcement) at
Lian-Yun-Gang, China, are presented.
Both the field measured data and the
results of finite element analysis are
presented. For the details of the FEM
analysis, readers can refer Chai and
Bergado (1993a) and Chai et al.
(2002).
Fig. 10 Geometry and instrumentation points of
A geogrid reinforced embankment Malaysian embankment
on Muar clay deposit, Malaysia

A geogrid reinforced embankment was constructed on soft Muar clay, Malaysia. The soil
profile at the site consisted of a weathered crust at the top 2.0 m which is underlain by about 5
m of every soft silty clay. Below this layer lies a 10 m thick layer of soft clay which in turn is
underlain by about 0.6 m of peat. Then, a thick deposit of medium dense to dense clayey-silty
sand is found below the peat layer. In order to ensure that the estimated settlement can be
obtained within the allowed construction period of 15 months, vertical band drain (without
filter) were installed in foundation with a square pattern of 2.0 m spacing to 20 m depth. After
vertical drain installation, geogrid were placed at the base of the embankment to ensure FS
about 1.3 during construction (Ting et al. 1989). Two layers of geogrid were laid in a 0.5 thick
sand blanket with 0.15 m vertical spacing. The strength of the geogrid was 110 kN/m with a
failure strain of 11.2%. A stiffness of 650 kN/m was adopted
The embankment was constructed with a base width of 88 m and length of 50 m, initially
to a fill thickness of 3.9 m. Then a 15 m wide
berm was left on both sides and the
embankment was constructed to a final fill
thickness of 8.5 m. The geometry and the field
instrumentation points are shown in Fig. 10.

Excess pore pressure


A typical variation of the excess pore
pressure with elapsed time for a piezometer
point 4.5 m below ground surface and on the
embankment centerline is shown in Fig. 11. In
the figure, Higher permeability means that
the vertical hydraulic conductivity is two time
of that Low permeability, and Varied Fig. 11 Excess pore pressure variation
(after Chai and Bergado 1993a)

- 77 -
permeability means that the initial values are the same as Higher permeability but varied
with void ratio following Eq. 6. It can be seen that the Varied permeability analysis,
simulated both excess pore pressure build up and dissipation for stage constructed
embankment well. The settlement and lateral displacement presented in the next paragraph are
the results of Varied permeability analysis.

Settlement and lateral displacement


The measured and simulated settlements and lateral displacement profiles are depicted in
Figs. 12 and 13. For this case, the finite element analysis simulated the field behavior very
well. Although this is a class C prediction, it is encourage that FEM has the ability to
predict the behavior of the reinforced embankment on soft subsoil.

Fig. 12 Comparing the settlement (after Fig. 13 Comparing the lateral


Chai and Bergado 1993a) displacement of Malaysian embankment
(after Chai and Bergado 1993a)
Effect of the reinforcement on stability
The calculated reinforcement tension
force distributions for different fill thickness
are shown in Fig. 14. It shows that at the
early stage, higher tension force developed
near the embankment toe at the location of
higher shear stress level zone. Later on,
since the berms are placed on both sides of
the embankment and also due to
consolidation effect, the tension force
increased at the embankment center position
and decreased under the berm. At the end of
the construction, the calculated maximum
tension force in each layer of the geogrids is
13 kN/m (totally 26 kN/m), which is Fig. 14 Simulated mobilized tensile force
equivalent to 2% of axial strain. Slip circle (after Chai and Bergado 1993a)
analysis result indicates that this tensile force can only increase the FS of the embankment
about 0.015. If the tensile strength can be fully mobilized, it can increase the FS about 0.12.

Soil/reinforcement interaction mode


Figure 15 shows the calculated shear stress distributions at soil/reinforcement interfaces of
bottom layer of the geogrid at different fill thickness. The sign convention is also shown in the
figure. It can be seen that the signs of shear stress at upper and lower interfaces are the same
for most interface areas, i.e. the direct shear interaction mode is applicable for this case. In the
zone near the toe of the embankment and the intersection point between the berm and the
main embankment, the lateral displacement of the fill is large because of the free face of the

- 78 -
embankment fill, and the interface shear stress has a negative sign. In other zones, the lateral
squeezing of the foundation soil causes the
interface shear stresses to have a positive sign.
If assuming the reinforcement as steel grid
with a stiffness of 156 MN/m, the pullout
interaction mode is observed.

Built to failure embankments at


Lian-Yun-Gang, China.

The test site is located in an alluvial plain,


Lian-Yun-Gang area, Jiangsu province, China.
The region belongs to lowland with an altitude
of 2.6 m. The ground water level fluctuated
between 0.5~1.0 m below ground surface. The
soil profile consists of 2.0 m thick clay crust
underlain by 8.5 m thick soft clay layer (called
mucky clay in China). Below the soft layer Fig. 15 Interface shear stress
there are medium to stiff sandy clay and silt distribution (after Chai and Bergado
sand layers. 1993a)
Two built to failure embankments
each had a length of 45 m. The CL 12.75 m 8.0 m
embankments had a base width of 42
Failure height (H) of embankment
m. After placing a 0.5 m thick sand Settlement
1:1
.7 5
on natural subsoil: 4.04 m

H
mat, a berm with a width of about 8.0 gauges
Geotextile
with reinforcement: 4.35 m

m was left on both sides and the

2.0 m

Mucky
embankments were built to failure. 2.0 m Sand mat 0.5 m

clay
The embankments had a 1V:1.75H
slope. Fill material was sandy clay and 2.5 m
Inclinometer
the average filling rate was about 0.1
Muck
casing
8.5 m
3.0 m
m/day. The embankment geometry, the
location of reinforcement, and the
Piezometers
main instrumentation points for
geotextile reinforced embankment are
illustrated in Fig. 16.
Sandy
clay

Two types of geotextiles were used.


One was a woven polypropylene
geotextile with a unit weight of 303 Fig. 16 Embankment geometry and main
2
g/m . In-air tensile strength from instrumentation points (after Chai et al. 2002)
wide-width strip test (ASTM 1994)
was 40 kN/m and failure strain was about 18% at a strain rate of 2%/min. Another was a
heat-bounded non-woven geotextile with a unit weight of 260.8 g/m2. In-air tensile strength
from wide-width strip test was 38.5 kN/m and failure strain was about 20%.

Effect of reinforcement on failure fill thickness


The field measured failure fill thickness was 4.35 m for reinforced case and 4.04 m for
unreinforced case. In analysis, the modeled strength of subsoil was adjusted (slightly change
the yield locus) in such a way that the analyzed failure fill thickness for unreinforced case is
the same as field value of 4.04 m. Then, the reinforced case was analyzed, which yielded a
failure fill thickness of 4.25 m and 0.10 m less than the field value.
In order to illustrate the effect of the reinforcement on embankment failure height, an
analysis with other conditions the same as reinforced case but no reinforcement (refer to as
assumed case) was conducted. The numerical results revealed that, for the assumed case, the

- 79 -
failure fill thickness is the same as reinforced case. The analysis indicates that the difference
of failure fill thickness of test embankments may not only be due to the effect of
reinforcement alone. The slight difference in construction schedule may also have contributed
to the difference of the embankment failure fill thickness, as more consolidation takes place
with longer construction period.
Since it is not convenient to obtain a value of FS from FEM analysis, the limit equilibrium
analysis was conducted to quantify the effect of reinforcement on FS of the embankment. The
result indicates that the reinforcement may only increase FS of the embankment by less than
0.055 (assume a mobilized tensile strength of 40 kN/m). Symmetric failure pattern was
observed in the field, i.e. the embankments failed at both sides. The observed failure surfaces
are not much different for with and without geotextile reinforcements. It supports the
hypothesis in conventional stability analysis that the existence of geotextile reinforcements
does not alter the failure surface much.

Surface settlement
The measured and simulated surface
settlement-time curves under embankment 4

Fill Thickness (m)


centerline together with embankment
construction histories are shown in Fig. 17. For
reinforced case, up to just before failure, the 2
FEM analysis fairly simulated the field value.
Field FEM
However, for unreinforced case, the analysis wirhout reinforcement
with reinforcement
over predicted the field value. Also, the 0
following observations can be made from the 0
figure. (a) Before fill thickness exceeded 2.0 m,
the settlement was small. The top crust is stiffer 0.1
and when the surcharge load was less than the
0.2
yield stress of the crust, the crust limited the Field
Settlement (m)

without reinforcement
subsoil deformation. (b) When the embankment 0.3 with reinforcement
approached failure-state, the settlement was FEM
without reinforcement
rapidly increased due to the lateral distortion of 0.4 with reinforcement
the subsoil. (c) Just before failure, measured
data shows that the geotextile reinforcement 0.5
0 10 20 30 40 50 60
slightly reduced the settlement rate. It is Elapsed Time (days)
reasoned that the combination of geotextile
(strong in tension) and top crust (strong in Fig. 17 Surface settlement under
compression) improved the contribution of the embankment centerline (after Chai et
top crust to the embankment stability. al. 2002)

Lateral displacement
0
The lateral displacement profiles of
reinforced case are given in Fig. 18. For -2
unreinforced case, the measured data are not Fill thickness:
-4 4.35 m (field)
available. It indicates that the lateral
Depth (m)

4.25 m (FEM)
displacement was small before the fill thickness -6 3.7 m
reached 4.35 m (failure occurred). Before the With reinforcement
-8 2.6 m
failure, the simulated values over-predict the Field
FEM
measured values and when the failure state is -10 Assumed without reinforcement
reached, the simulation under-predict the field FEM
-12
values at near ground surface. Generally, FEM 0 100 200 300 400
analysis yielded a poor simulation on lateral Lateral displacement (mm)
deformation. This is partially due to the adopted Fig. 18 Lateral displacement
soil model may not fully represent the behavior
profiles (after Chai et al. 2002)

- 80 -
of subsoil and partially due to the limitation of FEM analysis in simulating large deformation
at close to failure state. Comparing the assumed case (unreinforced) and reinforced case at the
same fill thickness, it can be seen that at lower fill thickness (2.6 m), there are very small
differences in the lateral displacement. However, when the fill thickness was increased to
approach the failure state, the lateral displacement of the assumed case is gradually becoming
larger than the reinforced case. Thus the confining effect of the reinforcement is apparent.

Excess pore
pressures
5
Variations of a) without reinforcement
Fill Thickness (m)

Fill Thickness (m)


4 b) Reinforced embankment
excess pore 4
3 FEM
pressure at 2.0 m, Field
Field
2 FEM
4.5 m (4.4 m for 2
unreinforced case), 1

and 7.5 m depths 0


80 0
Excess Pore Pressure (kPa)

80
under embankment

Excess Pore Pressure (kPa)


Depth Field FEM
centerline are given 60 2.0 m 60
Depth Field FEM
4.4 m 2.0 m
in Figs. 19 (a) and 7.5 m 4.5 m
(b) for unreinforced 40 7.5 m
40
and reinforced cases,
respectively. 20 20
Generally, the FEM
0
analyses simulated 0 10 20 30 40 50 0
0 10 20 30 40 50
the field data well. Elapsed Time (days) Elapsed time (days)
When fill thickness
was less than about Fig. 19 Excess pore pressure variations (after Chai et al. 2002)
2.5 m, the measured
data showed that during rest (consolidation) period between two load increments, there was a
tendency of dissipation of excess pore pressure due to consolidation effect. While when fill
thickness was more than about 2.5 m, even during rest period, field data showed that there
was no reduction or an increase in excess pore pressure at 4.5 m and 7.5 m depths. When fill
thickness approached 4.0 m, there was an obvious increase of excess pore pressure during rest
period. However, the FEM analysis failed to simulate this phenomenon because it did not
consider the creep behavior of the subsoil. It can be explained that due to the progressive
development of shear strain in the subsoil, shear induced excess pore pressure increment was
larger than partial dissipation effect. Also, when the subsoil approaches the failure state, the
coefficient of consolidation is reduced due to reduction of soil stiffness, which will reduce the
dissipation rate of excess pore pressure.

Mobilized tensile force in the reinforcement 40


Figure 20 depicts the calculated tensile forces H=4.3 m Just before embankment failure
Tensile force in geotextile (kN/m)

in the geotextile. Just before embankment failure,


30
for a stiffness J=1600kN/m case, the maximum
K=1600 kN/m
the tensile force is 36 kN/m, which is about 90% K=800 kN/m
20
of tensile strength. The corresponding mobilized H=3.7 m
tensile strain is 2.3%. For a stiffness J=800 kN/m
case, the mobilized maximum tensile force is 23 10
kN/m and about 60% of the tensile strength. The
corresponding mobilized tensile strain is 3%. We 0
0 5 10 15
consider that FEM analysis can only simulate Distance from the centerline of embankment (m)
pre-failure state, but not the failure process
accompanying a large deformation. In the field, Fig. 20 Simulated mobilized tensile
the rupture of the reinforcement was observed. force (after Chai et al. 2002)

- 81 -
However, we believe that the rupture might be a post failure phenomenon, i.e. it was ruptured
due to larger deformation of the embankment after the failure.

CONCLUSIONS

The mechanisms of reinforced embankment on soft subsoil as well as some modeling


techniques are discussed and two case histories are presented.
(1) Effect of the reinforcement on embankment behavior. The mobilized tensile force in the
basal reinforcement contributes to the stability of the embankment, and the confining
effect of the reinforcement to embankment fill material and soft subsoil can increase the
bearing capacity and reduce the lateral deformation of soft subsoil. However, the effect of
the reinforcement on subsoil deformation can only become significant when soft subsoil
approaches to failure.
(2) Some techniques for modeling the reinforced embankment on soft subsoil. (a)
Soil/reinforcement interaction modes are pullout and direct shear. In case of grid and strip
reinforcements, soil/reinforcement interface behavior is different for these two modes and
it should be properly modeled in numerical analysis. (b) During construction process, the
geometry of the embankment changes due to the deformation of soft subsoil. This is a
kind of large deformation problem and the numerical analysis should simulate it closely.
(c) Hydraulic conductivity of soft subsoil changes with the void ratio of soil and which
has to be considered in simulating the consolidation process.
(3) Two case histories of reinforced embankment on soft subsoil. One of the embankments is
a geogrid reinforced embankment on Muar clay deposit in Malaysian and another one is a
built-to-failure geotextile reinforced embankment in China. For both cases, the limit
equilibrium analysis results show that the mobilized reinforcement tensile force may only
had a marginal effect on FS of the embankment. For the built-to-failure embankment in
China, the analysis shows that only when the embankment approaches to failure, the
reinforcement has noticeable effect on lateral displacement of the soft subsoil. The
numerical analysis results also show that soil/reinforcement interaction mode is a function
of the relative stiffness of reinforcement and surrounding soil. In most cases, the direct
shear mode is a dominate one, and only when the reinforcement is very stiff, like steel
grid, the pullout mode becomes a dominate mode.

ACKNOWLEDGEMENT

This paper is part of research work conducted during the first authors sabbatical stay in
Tongji University, Shanghai, China.

REFERENCES

Asaoka, A., Nakano, M., and Matsuo, M. (1992). Prediction of the partially drained behavior
of soft clays under embankment loading. Soils and Foundations, JGS, 32: 41-58.
ASTM, D4595-86 (1994). Standard Test Method for Tensile Properties of Geotextile by the
Wide-width Strip Method. American Society of Testing and Materials, West
Conshohocken, Pennsylvania, U.S.A.
Bonaparte, R. and Christopher, B. R. (1987). Design and construction of reinforced
embankments over weak foundations. Transportation Research Record, 1153: 26-39.
Britto, A.M., and Gunn, M.J. (1987). Critical state soil Mechanics via finite elements. E11is
Horwood Ltd., Chichester, U.K.
Chai, J.-C. (2001). Linear ln(e+ec)-ln(pc) relation of structured natural clay. Proc. of 15th
Inter. Conf. on Soil Mech. and Geotech. Eng., Istanbul, 1: 59-62.
Chai, J.-C. and Bergado, D. T. (1993a). Performance of reinforced embankment on Muar
clay deposit. Soils and Foundations, JGS, 33(4): 1-17.

- 82 -
Chai, J.-C. and Bergado, D. T. (1993b). Some techniques for finite element analysis of
embankment on soft ground. Can. Geotech. J., 30: 710-719.
Chai, J.-C., Bergado, D.T., Miura, N., and Sakajo, S. (1996). Back calculated field effect of
vertical drain. Proc. Second Int. Conf. Soft Soil Engrg., Nanjing, China, 1: 270-275.
Chai, J.-C., Miura, N., Sakajo, S., and Bergado, D. T. (1995). Behavior of vertical drain
improved subsoil under embankment loading. Soils and Foundations, JGS, 35(4): 49-61.
Chai, J.-C. and Miura, N. (1999). Investigation of factors affecting vertical drain behavior.
Journal of Geotechnical and Geoenvironmental Engineering, ASCE, 125(3): 216-226.
Chai, J. C. and Miura, N. (2000). A design method for soft subsoil improvement with
prefabricated vertical drain. Proc. of Inter. Seminar on Geotechnics in Kochi, 161-166.
Chai, J.-C., Miura, N. and Shen, S.-L. (2002). Performance of embankments with and
without reinforcement on soft subsoil. Can. Geotech. J. 39: 838-848.
Davis, E. H. and Booker, J. R. (1975). Application of plasticity theory to foundation. In Soil
Mechanics recent developments, Edited by S. Walliapan, S. Hain, and I. K. Lee,
Butterworths, Sydney, N.S.W., Australia, Chapt. 3, pp. 83-112.
Hird, C. C. and Kwok, C. M. (1989). Finite element studies of interface behavior in
reinforced embankments on soft ground. Computer and Geotechnics, 8: 111-131.
Jewell, R. A. (1988). The mechanics of reinforced embankments on soft soils. Geotextiles
and Geomembranes, 7: 237-273.
Ladd, C. C. (1991). Stability evaluation during stage construction. J. of Geotechnical
Engineering, ASCE. 117(4): 541-615.
Milligan, V. and La Rochelle, P. (1984). Design methods for embankments over weak soils.
Proc. Symposium on Polymer Grid Reinforcement in Civil Engineering, Institute of Civil
Engineers, London, U. K.: 95-102.
Miura, N. and Chai, J.-C. (1999). A method for determining the in-soil strength and stiffness
of non-woven geotextile. Proc. of AIT 40th Anniversary Civil and Environmental
Engineering Conference, New Frontiers & Chanllenges, Bangkok, Thailand. 2: IV-29 to
IV-36.
Mylleville, B. L. J. and Rowe, R. K. (1988). Simplified undrained stability analysis for use
in the design of steel reinforced embankments on soft foundations, GEOT-3-88, Faculty
of Engineering Science, University of Western Ontario, 73p.
Rowe, R. K. and Li, A. L. (2002). Geosynthetic-reinforced embankments over soft
foundations. Proc. of 7th Inter. Conf. on Geosynthetics (editors Ph. Delmas and J. P. Gourc),
Balkema, Vol. 1: 5-34.
Rowe, R. K. and Soderman, K. L. (1985). An approximate method for estimating the
stability of geotextile-reinforced embankments. Can. Geotech. J., 22: 392-398.
Rowe, R. K. and Soderman, K. L. (1987). Stabilization of very soft soils using high strength
geosynthetics: the role of finite element analyses. Geotextile and Geomembrane, 6(1-3):
53-80.
Shen, S. L., Yang, C. W., Miura, N. and J. C. Chai (2000). Field performance of PVD
improved soft clay under embankment. Proc. of Coastal Geotech. Eng. in Practice,
edited by A. Nakase and T. Tsuchida, Balkema, 515-520.
Tavenas, F., Jean, P., Leblond, P.. and Lerouei1. S. (1983). The permeability of natural soft
clays. Part II: permeability characteristics. Can. Geotech. J., 20: 645-660.
Tavenas, F., Tremblay, M., Larouche, G., and Leroueil, S. (1986). In situ measurement of
permeability in soft clays. ASCE Special Conference on Use of In-situ Tests in
Geotechnical Engineering, Blacksburg, 1034-1048.
Taylor, D. W. (1948). Fundamentals of Soil Mechanics, John Wiley & Sons Inc. New York.
Ting, W. H., Chan, S. F. and Kassim, K. (1989). Embankment swith geogrids and vertical
drain. Proc. of Inter. Symp. on Trial Embankments on Malaysian Marine Clays, Kuala
Lumpur, 2: 21/1-21/13.

- 83 -

You might also like