Вы находитесь на странице: 1из 14

Computers and Geotechnics 28 (2001) 255268

www.elsevier.com/locate/compgeo

Analyses of stability of geogrid reinforced steep


slopes and retaining walls
M. Srbulov *
SAGE Engineering Ltd, Newark House, 2645 Cheltenham Street, Bath BA2 3EX, UK
Received 9 February 2000; received in revised form 15 November 2000; accepted 17 November 2000

Abstract
The results of measurements of axial strains in geogrids of two reinforced steep slopes and
two retaining walls were uniformly interpreted. The stabilities of slopes and walls are analyzed
using a method based on limit equilibrium. The method of analysis takes into account strains
along boundaries of rigid wedges in addition to the forces considered by classical methods of
limit equilibrium. However, the results obtained by the method remain only approximate due
to necessity to introduce a number of simplifying assumptions. # 2001 Elsevier Science Ltd.
All rights reserved.
Keywords: Stability analysis; Geogrid; Reinforced; Slopes; Retaining walls

1. Introduction
The use of natural bres to improve small tensile strength of in situ ground and
man made lls date to pre-historic times. However, the rational approach to the
problem of soil strength improvement by reinforcement is rather recent. In the late
fties, Henri Vidal, a French architect and engineer, launched a new civil engineering material known as Reinforced Earth. With rapid development of technology, the
initial metal strips are replaced by geotextiles and geogrids, which are almost exclusively used these days. Jewel [1] presented a rather comprehensive review of the
methods used for the design of reinforced soil.
The potential failure surfaces, most widely used for the analysis of internal equilibrium
of steep reinforced soil slopes and walls in connection with the limit equilibrium
* Tel.: +44-1225-486-500; fax: +44-1225-486-597.
E-mail address: milutin@sage- uk.com
0266-352X/01/$ - see front matter # 2001 Elsevier Science Ltd. All rights reserved.
PII: S0266-352X(00)00032-X

256

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

method, include the logarithmic spiral and two-part wedge mechanism. These two
failure mechanisms were found to be the simplest and at the same time the most
appropriate representatives of the true failure modes throughout the reinforced part.
These ndings will be re-examined in this paper for a number of cases of instrumented geogrid reinforced steep slopes and retaining walls.
2. Description of the extended method
Conventional limit equilibrium methods use a constant factor of safety FS along a
potential slip surface (and sometimes interfaces) under the assumption that soil
strength is mobilized at all places at the same (or similar) shear strains. Factor of
safety FS is usually dened as the ratio between available shear strength a and the
shear stress e necessary to maintain limit equilibrium
FS a =e

a is equal to the peak strength p when FS >1 or to the post-peak strength if


yielding occurs. Using MohrCoulomb failure criterion, which relates shear and
compressive stresses  (Fig. 1a), FS can be expressed in terms of these stresses
FS c tane

Fig. 1. (a) Shear strength versus compressive stress, (b) shear stress versus shear strain, (c) specic thickness change versus shear strain.

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

257

where c is cohesion and  angle of soil internal friction. Eq. (2) can be written in
terms of normal N and shear T force acting on a particular surface
FS cb Ntan=T

where b is width of the surface.


Knowing one of the components of the resultant forces acting along a slip surface
and a constant FS, it is possible to calculate the other components, so that the
number of unknown forces to be determined from available force and moment
equilibrium equations is decreased.
However, shear strains are seldom uniform in the eld even within a homogeneous
soil and at relatively small stress levels. When soil shear strength at large shear strain
is smaller than its peak value then localized and propagating (progressive) failure
may occur if induced shear strains are large enough. Similarly, if soil is heterogeneous and some parts mobilize and loose their peak shear strength while the other
parts are still on the way to mobilize the peak strength then again localized and
progressive failures may occur. The problem of analysing progressive failure may be
avoided if only soil residual strength is considered. Reinforced soil also experiences a
progressive failure. However, when soil reinforcement is loaded beyond its ultimate
strength it tends to break and completely loose all the strength as its residual
strength is zero. In the later case, the use of residual strength only is pointless
because it would lead to consideration of unreinforced soil.
Dierent methods have been proposed to solve the problem of propagating (progressive) failure using complete stressstrain solutions by nite and discrete/distinct
elements. This article describes a procedure for the consideration of the local and
progressive failures within the framework of the limit equilibrium method when
applied to analysis of stability of reinforced steep slopes and retaining walls. Srbulov
[2] used the method to analyze the progressive failures of actual but unreinforced
slopes in brittle soil. It has been shown that the extended method predicted well the
instability of the slopes, while a classical method provided FS much greater than 1
for the peak shear strength and much smaller than 1 for the residual shear strength
alone.
The activation of shear stresses a ; e is accompanied by development of shear
strains (Fig. 1b) and, therefore, shear stress/strength can be expressed as a function
of both compressive stress  and shear strain . The function can be determined
from soil shear tests. In its simplest form, when only the shape but not the value of
function  is assumed independent of  (Fig. 1b), the function becomes
 C  k

where C, k41 are soil constants determined by curve tting from laboratory test
results. This assumption can be considered reasonable for a rather large  stress
range only for cohesive soil under undrained conditions and reinforced soil when
reinforcement strength has a greater inuence on the behaviour of composite material. The  stress dependent shapes of the function may be introduced but on

258

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

account of the use of an additional iterative procedure that must be applied until the
dierences between initially assumed  stress levels are close to the calculated s
stress levels within desired tolerance. An alternative approach would be division of a
soil zone into subzones each corresponding to appropriate  stress level as it has
been done in the case when a nonlinear shear strength envelope is linearized within
chosen  stress intervals. Using Eq. (4), factor of safety at the surface i can be written in the form
k
Fi i;ak = i;e

k
is the shear strain corresponding to available shear stress a at a surface i
where i;a
k
and i;e is the shear strain corresponding to mobilized shear stress e at the surface i.
Similarly at another surface j,
k1
k1
Fj j;a
= j;e

where the exponent k1 is dierent from k in the case when dierent soil types exist at
the places i and j. From Eqs. (5) and (6) it follows
k1
k k
k1
= i;a
i;e = j;e
Fj Fi j;a

Unknown Fi can be determined from available equilibrium equations similarly to


a constant FS in conventional methods. Mobilized shear strains can be determined
indirectly from Eq. (6)
j;e j;a Fj

l=kl

kl
k
The values of j;a
and j;a
can be determined from soil shear strength tests. It should
k1
k
k
k1
be noted that the ratios j;a
= i;a
and i;e
= j;e
but not particular values of shear strains
are necessary to calculate the local factor of safety at any surface j. This has several
useful implications. If soil is tested in a direct shear apparatus, for example, then the
ratio of shear strains at available strength can be replaced by the ratio of measured
k
horizontal displacements k1
j;a =i;a instead. If soil shear tests are performed using a
triaxial apparatus then the results are presented as the deviatoric stress versus axial
strain". If the properties of all soil types involved in an analysis are measured and
presented consistently in terms of axial strain than it is possible to use the ratio
k
k1
k
"k1
j;a ="i;a in place of the ratio j;a = i;a The reinforcement strength is usually described
as a function of axial strain.
If a soil zone is divided into wedges and the wedges between their boundaries do
not change their volumes (they are rigid) except that some local yielding is allowed
k
k1
at their tips then the ratio j;e
= i;e
between the magnitudes of actual shear strains at
two boundaries (i, j) will be the same to the ratio ki;e =k1
j;e of magnitudes of potential
shear strains along these boundaries, because and  will be directly proportional
( B, where B is an unknown constant). For rigid wedges, tangential displacements along a particular boundary will be constant. In undrained conditions, there

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

259

will be no volumetric changes along wedge boundaries and potential shear strain can
be dened as the ratio between potential tangential displacement along the boundary and the boundary thickness. For a unit thickness, the shear strains will be equal
to the tangential displacements. Potential tangential displacements, illustrated in
Fig. 2 for the case of zero soil volume change in undrained conditions, can be
dened starting with a unit tangential displacement along the base of the rst wedge.
Proceeding along the external boundaries so that the known tangential displacement
is the vectorial sum of the displacement in the direction of the interface and base of
the following wedge it is possible to dene all other potential displacements.
The assumption of no soil volume change along a boundary is not correct for
drained conditions nor is it correct to assume that the wedges are rigid. However,
the volumetric changes due to axial stresses should be small for sti soil and at
relatively small  stresses acting within steep slopes. When they are likely to be signicant such in soft clay and loose sand then the assumption of no volumetric strain
due to axial stresses in drained condition is less acceptable.
Volumetric strain "v (i.e. specic thickness dt) change along the boundaries with
shear strain change can be taken into account in the construction of strain diagram
in Fig. 2 as the inclinations of strain vectors with respect to the boundaries. From
calculated Fj using Eq. (7) it is possible to back calculate j;e using Eq. (8). The
function of volumetric strain (specic thickness dt) change versus shear strain
(Fig. 1c) can be determined from soil shear tests. The angle of inclination j of shear
strain vector with respect to the boundary is simply

j arctan dtj;e = j;e
9
If local over stressing occurs at face j (which means that Fj tends to become less
than 1 and the mobilized strength greater than the peak value, which is impossible)

Fig. 2. Possible tangential displacements 0 s along boundaries of wedges.

260

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

then the yielding, accompanied by an increase in shear strains j;e can be simulated
by the increase of j;a in Eq. (7) until Fj becomes equal to 1. Such simulation is
necessary because xed value j;e , dependent on geometry only, is used instead of
actual shear strain j;e . With post peak increase (yielding) of shear strain j;e , brittle
soil will soften and, therefore, the corresponding decrease in the shear strength
parameters c, with the increase in j;e is taken into account in Eq. (3). A change
(increase) in a factor of safety Fj will cause a corresponding change in Tj force and
all other forces because of the need to satisfy the equilibrium equations. The eect of
the rate of soil shear strength decrease after the peak value has been investigated for
two simple wedges before [3]. It has been found that there will be an evident dierence in the calculated factors of safety of the slope stability if soil decreases its peak
strength towards the residual value more gradually then abruptly. In the former case
the use of an instant jump from the peak to the residual strength is not appropriate.
The calculated factors of safety by the extended method and by a classical method
were also in a good agreement if soil exhibit an abrupt shear strength drop after the
peak. However, the classical method yields smaller factors of safety than the extended method for gradually decreasing post peak shear because the mobilized shear
strength is somewhere between its peak and the residual value. Two wedges, non
brittle soil with shear strength parameters c=0, 0 24 and soil unit weight of 22
kN/m3 are used for an investigation of the eect of change in volumetric strain
(specic thickness) with shearing. A function describing the specic thickness change
with shear strain (Fig. 1c) is adopted in the form which is appropriate for both the
contracting and dilating phases of soil volume changes up to its constant value at the
steady-state
dt 1=2 a ln b

10

where a,b are constants or functions of axial stress which can be determined by tting experimental data. The results of computations are shown in Table 1.
The eect of volumetric strain changes with shearing is not very important for the
majority of slightly dilatant/contractant soil but could be very important for heavily
overconsolidated clay and very dense sand.
Table 2 contains the list of unknown values, available equations and their numbers for n wedges.
It can be noted that the use of the local factors of safety Fj increased the number
of unknown values for 2n 2 as well as that the number of available equation types
[7] increased for 2n 2 in comparison with the number of unknown values and
available equations in conventional methods of limit equilibrium. It is also evident
Table 1
The eects on factor of safety of soil dilatation ( ) and contraction (+) with shearing of soil along
boundaries of the two wedges
Peak volume/peak shear strain ratio
Factor of safety

0.6

0.4

0.2

3.17

1.80

1.73

+0.2

+0.4

+0.6

1.67

1.63

1.59

1.51

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

261

Table 2
Values, available equations, and their numbers for n wedges
Unknown values

Number

Available equations

Number

Normal forces N at bases


Location of N at the last base
(at other bases assumed in
the middle of bases)
Normal forces N at interfaces
Locations of N at interfaces
Shear forces T at bases
Shear forces T at interfaces
Local factors of safety Fj
at bases (except at i)
at interfaces

n
l

Forces equilibrium in horizontal direction


Forces equilibrium in vertical direction
Moments equilibrium

n
n
n

n 1
n 1
n
n 1

k1
k
k1
= i;a
ki;e = j;e
:
Fj Fi j;a
at bases (except at i)
at interfaces

Tj cj bj Nj tanj =Fj :
at bases
at interfaces

n 1

Total

7n 3

Total

7n 3

n 1
n 1

n 1
n 1
n

that the positions of normal forces N's at the bases (except at the last base) are
assumed to be in the middle of bases. Various assumptions have to be introduced in
all procedures based on limit equilibrium method due to excessive number of
unknown values in comparison with available limit equilibrium equations. Such
assumptions cause that the solutions obtained by the methods are only approximate
ones and not necessarily correct with regard to other stress-strain constitutive laws.
The system of 3n equilibrium equations is nonlinear due to unknown Fi in the
denominators of the coecients of equations. It is possible to apply an iterative
procedure by choosing an initial Fi(=1), solve 3n 1 linear equations, check the 3nth
equation and gradually change (increasing) Fi in steps until all 3n equilibrium
equations are satised to a specied tolerance. Several iterations will be necessary
for each step if local yielding occurs and therefore the coecients of the equations
must be readjusted. For an unstable wedge assembly, the equilibrium equations
cannot be satised and the stepping procedure will continue until permitted by the
user. For a stable wedge group, an average factor of safety of the group stability
Favr . can be calculated from the formula


Favr:  a;j bj = a;j bj =Fj
11
where j=1...n; Favr is used for comparison with a constant FS from conventional
methods and for the assessment of global stability of a group of wedges.
The expression relating forces and shear strains follows from Eqs. (3) and (7)

cj bj Nj tanj
Tj
12
k
Fi k1 j;a = i;a
ki;e =k1
j;e


k1
k
All local Fsj will be the same if soil is homogeneous j;a
i;a
and the ratios
ki;e =k1
j;e 1 at the bases of wedges along a plane or a circular cylinder with the
interfaces passing through the centre of the circle.

262

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

When only one layer of reinforcement exists, such as in reinforced embankments,


then reinforcement can be treated as a thin soil layer, Srbulov [4]. However, when
many reinforced layers are present it is necessary to apply a composite material
concept. Ingold [5] referred to the work of Long et al. [6] who observed that above a
certain threshold value of applied conning pressure in triaxial apparatus there was
a constant increase in applied vertical stress at failure in samples with reinforcement
at a given tensile strength and spacing, like in Fig. 1a. Failure of the reinforced
samples was very brittle (like in Fig. 1b), with a drastic decrease in strength when the
peak was passed. The brittleness was less severe at higher applied conning pressures or in less heavily reinforced samples. Post-failure inspection of dismantled
samples consistently showed that the reinforcement had failed in tension. It was
concluded that since, for tensile reinforcement failure, the failure envelopes of both
the reinforced and unreinforced sand are parallel, and therefore exhibit the same
angle of internal shearing resistance, the additional strength imparted by the reinforcement could be represented by an apparent cohesion.
In addition, the orientation of reinforcement is likely to change with respect to a
wedge boundary, from as-built position to an almost parallel orientation at the
failure. Jewel [7], proposed that the improvement in shearing resistance Ps, resulting
from a reinforcement force, Pr, can be expressed by the equilibrium equation of
forces
Ps Pr sin costan0

13

where  is the angle between the reinforcement direction and a normal to the
boundary. Eq. (13) also denes the degree of anisotropy of shearing resistance of
reinforced soil with respect to the angle . For 0 close to 30 it follows from the
above equation that almost a constant tensile force greater or equal to Pr acts for the
range of  between 30 and 90 . In the examples which follow, it will be assumed that
an apparent cohesion is given by the expression
c Pr =b

14

where  is applied to all reinforced layers crossing a particular considered wedge


boundary if the distance to the end of reinforcement is sucient for activation of
reinforcement tensile strength. It also means that reinforced soil shear strength is
considered isotropic, independent of the angle .
3. Case histories
3.1. Case 1.1large scale model of a reinforced slope
Kutara et al. [8] described a large-scale model of six meter high instrumented
embankment, with the slope inclination of 1 vertical to 1 horizontal, made of compacted silty sand, reinforced by geogrid. The embankment was submerged and the

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

263

water level was rapidly lowered in order to estimate the embankment stability using
a rotational slip surface and limit equilibrium method.
The embankment sand contained 30% of silt. The shear strength parameters
obtained by triaxial compression test and used in the stability analysis by the
authors were c0 =10 kPa and 0 32 . However, the test results have not been presented in the paper and the cohesion intercept may be a result of the envelope curvature or the way of interpretation and will not be used for the back analysis. The
dry unit weight was 13 kN/m3 and the relative density of 87%.
The reinforcement used was Tensar SS2 biaxially stretched polymer grid, placed in
four 6 m long layers at spacing of 1.5 m. The tensile strength specied in the paper is
15 kN/m at the peak axial strain of 5%. The slope surface was encapsulated with the
grid.
The behaviour of the embankment and the geogrid was monitored by strain gages
attached to the geogrids at interval of 0.5 m, by measurement of the settlement of
the embankment crest and its interior, by the measurement of horizontal and vertical strains of the slope and by the measurement of water level inside the embankment.
In addition a surcharge was placed on top of the embankment. The surcharge was
placed at a distance of 4 m from the slope crest. A 3 m high ll made of river sand
was used as a surcharge of 43.5 kN/m2.
The submergence of the embankment was achieved by water up to the height of
5.25 m. The water level had been rapidly lowered with the rate of 0.75 m/h in front
of the slope after the deformation and the strain of the geogrid ceased to increase
through several days of submergence.
The measured geogrid strain and embankment settlement at the end of rapid
water level decrease in front of embankment are shown in Fig. 3. The authors also
presented the strain diagrams for the construction phase, submergence and dierent
water level fall cases. Their shapes generally coincide with the nal shapes of the
strain diagrams. The face of slope also bulged greatly between the geogrid layers up
to the maximum of 38 cm.

Fig. 3. Case 1.1 geometry, ground water level, measured geogrid strain and embankment settlement.

264

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

Using a rotational slip surface and the equilibrium of moments, the authors
obtained the safety factor of 1.62 at the end of embankment construction, 1.32 at
the end of surcharge addition, 1.38 at the end of submergence and 0.97 at the end of
sudden water drawdown. However, they did not give details about the method of
stability analysis and the analyzed critical slip surface.
For the stability analysis by the method presented in this paper, the location of a
potential slip surface and the interfaces are determined from the positions of the
peaks in the geogrid strain diagrams and the shape of top settlement diagram shown
in Fig. 3. An almost rotational slip surface shape with inclined interfaces was inferred from Fig. 3. The value of axial strain at failure of 5% and the exponent k=0.5
are adopted for silty sand based on experience. The eect of the assumptions made
appeared to be of no great inuence on the results. The exponent k=0.5 for the
composite material (reinforced soil) was adopted based on the tensile strength-strain
diagram shown in the paper. The apparent cohesion used for the sides No. 1, 5, 6
was 0 kPa, No. 2, 4 was 10 kPa, No. 3, 11 was 5 kPa, No. 7, 8, 10 was 4 kPa and No
9 was 8.5 kPa. The unit weight of soil used was 19 kN/m3. Calculated Favr =1.04 by
the extended method is close to one and the real slope state. The local factors of
safety along the boundaries 16 are all equal to 1 and the average factor slightly
greater than one is the result of the local factors of safety along the interfaces
(Fs,7=1.5, Fs,9=1.04, F1,11=1.09). Mobilized axial strains are calculated according
to the equation
"j;e "j;a Fi

1=k1

15

At the boundaries 14 they are all 17.3% and at the boundaries 56 they are
19.4%. However, the corresponding axial displacements can not be calculated due
to the lack of a referent length.
3.2. Case 1.2large scale model of a steep reinforced slope
This case was described by the same authors as Case 1.1. The only dierence was
that the geogrid spacing was 0.75 m and that seven layers were used instead of four.
As a result of denser reinforcement, the maximum bulging of the embankment slope
between the reinforcement after rapid water level draw-down was 25 cm instead of
38 cm and the top settlement was more uniform. Also, the authors of the paper
reported the safety factor of 1.7 at the end of embankment construction, 1.38 at the
end of surcharge addition, 1.44 at the end of submergence and 1.01 at the end of
sudden water draw-down. They gave no details about the rotational slip surface and
the method used for the calculation of factor of safety.
The position of potential slip surface and the interfaces was inferred from the
peaks of measured geogrid strain diagrams and the shape of top settlements shown
in Fig. 4. From Fig. 4, it can be seen that the geogrid strain is sometimes uniform
over a considerable length. This is the result of superposition of strain when several
slip surfaces cross the geogrid at rather close distances. The apparent cohesion used
for the sides No. 1, 6, 7, 8 was 0 kPa, No. 2, 5, 9, 15 was 15 kPa, No. 3, 4, 10, 14 was

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

265

Fig. 4. Case 1.2 geometry, ground water level, measured geogrid strain and embankment settlement.

13 kPa, No. 11, 13 was 11kPa and No. 12 was 15kPa. The extended method failed to
satisfy equilibrium conditions.
3.3. Case 2large scale model of a reinforced retaining wall
The same authors (Kutara et al., [8]) investigated and presented the results of
measurements and stability analyses for a retaining wall 6 m high, with slope
inclined 1 vert. to 0.2 hor. with precast concrete panels over the surface. The panels
were joined to the reinforcement by reinforcing bars on the rear sides. The ll soil
was silty sand as for the slopes.
The reinforcement used was Tensar SR55 uniaxially stretched polymer grid,
placed in six 4 m long layers at spacing of 1 m. The tensile strength specied in the
paper is 55 kN/m at the peak axial strain of 14%. To strengthen the wall face, the
reinforcement was interposed by reinforcing material of the same type with length of
1 m.
The results of measured geogrid strain and the surface strains are shown in Fig. 5.
The authors also presented the strain diagrams for the construction phase, submergence and dierent water level fall cases. Their shapes generally coincide with
the nal shapes of the strain diagrams. For the back analysis, the location of a
potential slip surface and the interfaces are determined from the positions of the
peaks in the geogrid strain diagrams and the shapes of surface displacements diagram shown in Fig. 5.
The authors used a two-wedge failure mechanism with a vertical smooth interface
between the wedges for the analysis of wall stability by limit equilibrium method.
They calculated the safety factor at the end of sudden water decrease FS=1.9 if the
ll cohesion was taken into account and less than 1 if it was ignored using the
equilibrium of horizontal forces only. When the failure mechanism shown in Fig. 5
by thick dashed line is considered and silty sand cohesion is ignored then the exten-

266

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

Fig. 5. Case 2 geometry, ground water level, measured geogrid strain and wall displacements.

ded method gave Fave =1.36. More detailed results of the latter analyses are shown
in Table 3. From this table it can be seen that the local factors of safety at the failure
surface are all equal to 1 and that the Favr >1 is the result of larger factors of safety
along the interfaces.
3.4. Case 3large scale model of a reinforced retaining wall
Thamm et al. [9] described the results of measurements and calculations performed for a 3.6 m high retaining wall loaded by top surcharge to failure. The soil
used for the ll was a gravely sand with density of 1.95 g/cm3 and the angle of
internal friction 0 =39 determined from direct shear tests within the range of normal stresses from 40 to 120 kPa. The strain at failure of 2% and the exponent k=0.5
for gravelly sand were adopted based on the experience.
The reinforcement was combined of two dierent geogrids. The main reinforcement was made of Tensar SR2 geogrid with the peak strength of 67 kN/m at the
axial strain of 11%. The geogrid was 2.7 m long, placed in three bottom layers at the
Table 3
Detailed results of the analyses
No.

c0 (kPa)

N (kN)

T (kN)

M/Na (m)

T/N

Fs,1

"e (%)

1
2
3
4
5
6
7

0
40
35
0
40
28
40

17.5
153
69
251
80
134
169

7.5
51
30.6
155.5
50
61
45.5

0
0
0
0.08
1.73
2.36
0.78

0.43
0.33
0.44
0.62
0.62
0.46
0.27

1
1
1
1
1.96
1
3.06

64.4
19.6
19.6
47.6
3.6
14
1.5

M/N is the eccentricity of the resultant force acting on a surface.

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

267

spacing of 0.8 m and the two top layers at the spacing of 1 m. The additional reinforcement was 0.5 m long Tensar SS2 grid (with the peak strength of 31,5 kN/m at
the strain of 10.5%) near the wall surface, which was cascaded with the steps 0.2 m
at each geogrid level. These two geogrids were connected by a HDPE rope. TERRAM 2000 nonwoven geotextile was placed along the wall face to prevent the loss of
soil from the part of the wall.
The loading was applied over a reinforced concrete slab (2.40.9 m) placed 0.5 m
apart from the wall top edge. The authors used two-wedge sliding mechanisms with
a vertical interface and a slip circle surface to calculate failure loads of 950, 620, and
1296 kN respectively. The load was applied in steps and the nal failure load that
caused excessive settlements of the concrete slab and after some time the nal failure
was 1065 kN. For the load just prior the failure of 1026 kN, the geogrid strain
reached the maximum value of 2.9% in the top reinforcement. For the lower two
geogrid layers the maximum measured strains were about 1.7% (layer three) and
0.9% (layer two). The failure strain reached was 5%. The shapes of the geogrid
strain at dierent load steps were very similar.
The graphs of geogrid strain at the load level of 1026 kN together with the wall
horizontal strains are shown in Fig. 6. The position of potential failure surface was
inferred from the peaks in the strain diagrams and coincides well with the measured
pattern of wall strain. The position of the interface is adopted under assumptions
that the interface would not cross the reinforcement, which would provide greater
resistance to shearing and prevent the establishment of the minimum potential
energy at failure. The apparent cohesion for the sides No. 1 and 2 was 67 kPa while
for all the other sides 0 kPa. For the load of 1026 kN, the extended method indicated failure of the wall.

Fig. 6. Case 3 geometry, load, measured geogrid strain and wall displacements.

268

M. Srbulov / Computers and Geotechnics 28 (2001) 255268

4. Conclusion
From preceding examples it is evident that the positions of the peak values of
strain in the geogrids are aligned along distinct and almost planar interfaces. From
Figs. 36 it can be seen that the smaller spacing between the reinforcement provides
smoother slip surfaces and that typical shapes for the steep slopes were circular
while for the retaining walls were ``y'' shaped wedges with inclined interfaces.
The extended method with introduced force-strain relationships could provide
more realistic solutions than classical limit equilibrium methods, which cannot
account for potential progressive failure in brittle soil-geogrid composites. These
solutions remain only approximate due to the need of introduction of various
assumptions such as that the resultant forces along bases of wedges except the last
base act in the middle of the bases, that volumetric strains caused by axial stresses
are zero and that the shape of shear stressstrain function is independent of the axial
stress level. This later assumption can be avoided on account of introduction of an
additional iterative procedure or the eect of the assumption can be decreased if
axial stress level dependent zonation is introduced.
The method is simple to use and requires less input data than complete stress
strain methods. It inherited the same disadvantages of conventional limit equilibrium methods that a sliding mechanism must be assumed to govern slope stability.
References
[1] Jewel RA. Soil reinforcement with geotextiles. CIRIA Special Publication 1996;123:456.
[2] Srbulov M. A simple method for the analysis of stability of slopes in brittle soil. Soils and Foundations 1995;4:1237.
[3] Srbulov M. On the inuence of soil strength brittleness and nonlinearity on slope stability. Computers and Geotechnics 1997;20(1):95104.
[4] Srbulov M. Force-strain compatibility for reinforced embankments over soft clay. Journal Geotextiles and Geomembranes 1999;17(3):14756.
[5] Ingold TS. Reinforced Earth. London: Thomas Telford Ltd, 1982 910.
[6] Long NT, Guegan Y, Legeay G. Etude de la terre armee a l'appareil triaxial. Rapport de Recherche
1972;17:LCPC 6.
[7] Jewel, R.A., Strength and deformation in reinforced soil design. In: Hoedt D., editor. Proc. 4th
International Conference on Geotextiles, Geomembranes and Related Products. The Hague, Netherlands, Den Hoedt (Ed.), Balkema, Rotterdam, 1990, Vol. 3, 913946.
[8] Kutara K, Miki H, Kudoh K, Nakamura K, Minami T, Iwasaki K, Nishimura J, Fukuda N, Taki
M. Experimental study on prototype polymer grid reinforced retaining wall. In: Hoedt, D. (Ed.),
Proc. 4th International Conference on Geotextiles, Geomembranes and Related Products. The
Hague, Netherlands, Balkema, Rotterdam, 1990, 1, 7378.
[9] Thamm BR, Krieger B,. Krieger J. Full-scale test on a geotextile reinforced retaining structure. In:
Hoedt, D. (Ed.), Proc. 4th International Conference on Geotextiles, Geomembranes and Related
Products. The Hague, Netherlands, Balkema, Rotterdam, 1990, Vol. 1, 38.

Вам также может понравиться