You are on page 1of 11

Computers and Geotechnics 77 (2016) 45–55

Contents lists available at ScienceDirect

Computers and Geotechnics


journal homepage: www.elsevier.com/locate/compgeo

Research Paper

Probabilistic stability analysis of simple reinforced slopes by finite


element method
Ning Luo a,1, Richard J. Bathurst b,⇑, Sina Javankhoshdel a,1
a
GeoEngineering Centre at Queen’s-RMC, Department of Civil Engineering, Ellis Hall, Queen’s University, Kingston, Ontario K7L 3N6, Canada
b
GeoEngineering Centre at Queen’s-RMC, Department of Civil Engineering, 13 General Crerar, Sawyer Building, Room 3417, Royal Military College of Canada, Kingston, Ontario
K7K 7B4, Canada

a r t i c l e i n f o a b s t r a c t

Article history: Probabilistic slope stability analyses of simple geosynthetic reinforced soil slopes were carried out using
Received 17 November 2015 the shear strength reduction method in combination with the finite element method (FEM). An existing
Received in revised form 31 March 2016 open-source FEM code was modified to include bar elements to model horizontal layers of geosynthetic
Accepted 1 April 2016
reinforcement. Analysis results using the modified code (mFEM) demonstrate that large reductions in
probability of failure can be realized by adding geosynthetic reinforcement layers to constructed slopes.
The modified code was also used to investigate the effect of variability of soil friction angle on
Keywords:
probabilistic outcomes for constructed unreinforced and reinforced purely frictional soil slopes.
Probabilistic analysis
Reinforced slope stability
Crown Copyright Ó 2016 Published by Elsevier Ltd. All rights reserved.
Finite element method
Monte Carlo simulation
Limit equilibrium method

1. Introduction [33,11,15], amongst many others), the influence of soil material


variability on margins of safety expressed in probabilistic terms
Geosynthetic-reinforced slopes (and embankments) are widely for reinforced slopes and embankments has received little
used in geotechnical engineering. Since the early 1980s, conven- attention.
tional deterministic limit equilibrium methods (LEMs) for unrein- Kitch [26] carried out probabilistic analyses of two reinforced
forced slopes have been modified to include the stabilizing force slope examples with reinforcement layouts initially selected using
(or moment) contribution of geosynthetic reinforcement layers in design charts based on deterministic limit equilibrium methods.
constructed slopes and embankments. These methods include cir- Low and Tang [30] proposed a limit equilibrium stability model
cular slip [26], log-spiral [29] and two-part wedge [34,8] for reinforced embankments on soft ground and a practical reliabil-
approaches. The margin of safety is computed as a single-value ity evaluation procedure. However, the LEM approach used in both
critical factor of safety for the reinforced slope. studies has the disadvantage that the type of critical failure surface
It has been frequently demonstrated in the literature that two must be assumed a priori (e.g. circular, non-circular or bi-linear)
nominally identical natural slopes can have the same factor of and an assumption must be made regarding the magnitude and
safety based on conventional deterministic factor of safety analysis distribution of available stabilizing reinforcement tensile forces.
methods but have very different probabilities of failure due to ran- More recently, a probabilistic analysis technique called the
dom and spatial variability of the soil properties [38,13,19,9,22]. Random Finite Element Method (RFEM) has been developed by
While uncertainty in the properties of engineered fills used in Griffiths and Fenton [19] to conduct probabilistic stability analysis
constructed unreinforced and reinforced slopes are likely low of slopes with spatial variability of soil properties based on random
compared to values reported for natural in situ soil materials (e.g. field theory. In the limit of infinite spatial correlation length the
Random Finite Element Method can be understood to be the FEM
Abbreviations: FEM, finite element method; LEM, limit equilibrium method; with only random variability of soil properties. Conventional deter-
mFEM, modified finite element method; pLEM, probabilistic limit equilibrium ministic FEM slope stability analyses are also possible by assigning
method; RFEM, random finite element method. only constant values for the soil parameters. The shear strength
⇑ Corresponding author. Tel.: +1 (613) 541 6000x6479/6345/6391.
reduction method is used to compute the factor of safety in each
E-mail addresses: ning.luo@queensu.ca (N. Luo), bathurst-r@rmc.ca
(R.J. Bathurst), s.javan.khoshdel@queensu.ca (S. Javankhoshdel).
analysis. An advantage of the combined FEM-shear strength reduc-
1
Tel.: +1 (613) 541 6000x6345. tion method is that this approach allows the program to search out

http://dx.doi.org/10.1016/j.compgeo.2016.04.001
0266-352X/Crown Copyright Ó 2016 Published by Elsevier Ltd. All rights reserved.
46 N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55

the critical failure mechanism without constraints imposed by are taken from probabilistic stability design charts for simple unre-
fixed failure geometry assumptions [20]. inforced purely cohesive soil slopes developed by Javankhoshdel
In the current study an existing open-source FEM program and Bathurst [22]. They expressed Taylor’s slope stability equation
described by Griffiths and Fenton [19] and Fenton and Griffiths [39] using random variable notation as follows:
[15] was expanded to allow for probabilistic analysis of reinforced lsu
slope cases. Although the original and expanded code can consider Fs ¼ ð1Þ
lc HNs
spatial variability of soil properties, in this investigation only ran-
dom soil variables were considered in the probabilistic analyses.
Here F s is the mean factor of safety computed using mean values of
Hereafter, the original code is referred to as the FEM code or
Su and c (lsu and lc), slope height H and slope stability number Ns.
program and the modified code is referred to as the mFEM code
The value of F s from Eq. (1) together with coefficients of variation of
or program to avoid confusion with the random field theory
undrained shear strength (COVsu) and unit weight (COVc) were used
capability of the original source program.
to calculate probability of failure, Pf as follows [22]:
The original (unmodified) FEM code was first validated by com-
8 rffiffiffiffiffiffiffiffiffiffiffiffiffiffi  9
paring results of conventional and probabilistic analysis of unrein- > 1þCOV2su >
forced slope cases. In this paper, probabilistic slope stability
>
< ln 2 =F s
>
=
1þCOVc
analyses using LEM methods together with probability theory are Pf ¼ p½F s < 1 ¼ U rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
h     i ð2Þ
>
> >
referred to as probabilistic limit equilibrium methods (pLEMs). : ln 1 þ COV2su 1 þ COV2c > ;
Deterministic analysis results using the (modified) mFEM source
code (i.e. without random soil properties) were also compared to where U is the cumulative standard normal distribution function.
results of FEM analysis of reinforced slopes using a commercial Fig. 1a and b present two sets of results. The first considers the unit
FEM software package and assuming purely frictional soils. Good weight c to be deterministic (no variability) and hence COVc is
agreement was demonstrated between programs for the prediction equal to zero. For the second set, both undrained shear strength
of reinforcement strains giving confidence that implementation of Su and unit weight c are considered as uncorrelated lognormal
the reinforcement capability in the mFEM code is correct. Predic- distributed random variables. Therefore the coefficient of variation
tions of factor of safety and failure modes of reinforced slopes of factor of safety is due to the variability in random variables Su and
using the mFEM and LEM (Bishop’s Simplified Method with addi- c, and is calculated as:
tion of reinforcement forces) were also carried out. Differences in qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
results were attributed to the treatment of stabilizing forces in COVFs ¼ COV2su þ COV2c ð3Þ
LEM analyses. A strategy proposed by Hammah et al. [21] is used
to eliminate discrepancies between LEM and the coupled FEM- The dashed curves in Fig. 1a and b are the closed-form solutions
strength reduction method for the calculation of factor of safety using Eq. (2). The data show that as the mean factor of safety
for simple homogenous soil slopes. increases for any constant level of variability in random variables,
The utility of the expanded mFEM code for probabilistic analysis the probability of failure decreases, which is expected. Unreason-
of the factor of safety for simple soil slopes with frictional soils is ably large COV values appear in these figures. They were purposely
demonstrated using a number of unreinforced and reinforced slope used to test the accuracy of the FEM numerical code over a wide
examples. In the probabilistic study, the influence of variability of range of input values.
soil friction angle (/) and unit weight (c) on the probabilistic out- Fig. 2 shows the simple slope geometry used in the simulations.
comes of both unreinforced and reinforced purely frictional soil In order to compare unreinforced slope FEM outcomes with pLEM
slopes is investigated. The effect of variability in c and cross- results, two different groups of probabilistic analysis were con-
correlation between / and c on probability of failure is investigated ducted. In the first group the undrained shear strength Su was the
and (as expected) shown to have no effect. only random variable, while in the second group both the
undrained shear strength Su and the unit weight c of the soil were
treated as random variables. The mean value of the unit weight
2. Verification of FEM code for unreinforced slopes was 20 kN/m3. All random variables are assumed to have lognormal
distributions. For deterministic stability analysis for the ultimate
2.1. General failure limit state, the choice of Young’s Modulus (E) and Poisson’s
ratio (m) has little influence on stability analysis outcomes [20],
With the exception of modifications for reinforced slopes, the hence parameters E and m were taken as 100 MPa and 0.3, respec-
open-source FEM code (mrslope2d) described by Fenton and Grif- tively. The mean value of Su was varied from 30 to 60 kPa with an
fiths [15] and available at ‘‘http://courses.engmath.dal.ca/rfem/” increment of 5 kPa. For the first group of analyses the undrained
was used in the current study. The deterministic slope stability shear strength Su was the only random variable and it was deter-
analysis part of the code (Program 6.4, [37] is based on the shear mined that 1000 Monte Carlo simulations were sufficient to give
strength reduction method. This program is for two-dimensional a consistent estimate of probability of failure using FEM. However,
slope stability analysis of unreinforced slopes with elastic- for the second group with two random variables, 2000 Monte Carlo
perfectly plastic soils governed by the Mohr–Coulomb failure crite- simulations were needed to obtain a consistent result.
rion. Eight-node quadrilateral elements are used. The bottom of the The simulation results for the first and second groups of simu-
slope foundation is fixed in both horizontal and vertical directions. lations are plotted as symbols in Fig. 1a and b, respectively. The
The vertical boundaries on both sides are fixed in the horizontal pLEM results based on the closed-form solution (Eq. (2)) and
direction. The gravity ‘‘turn-on method” is used in Program 6.4 FEM outcomes can be seen to agree very well.
(see [20]).
2.2.2. Cohesive–frictional (c–/) soil slopes
2.2. Comparison of FEM with pLEM approaches (unreinforced slopes) Javankhoshdel and Bathurst [22] also developed probabilistic
slope stability charts for simple unreinforced cohesive–frictional
2.2.1. Cohesive soil slopes (c–/) soil slopes using pLEM (LEM with Monte Carlo simulation).
Results using the FEM method (i.e. original code without In these charts both cohesion c and friction angle / were
reinforcement) are first compared to pLEM results where the latter considered as random variables having lognormal distributions.
N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55 47

(a) 100
100 FEM
COVγ = 0 90 pLEM
90
80 a. COVc = 0.1, COVφ = 0.1
80 COVsu = 8 b. COVc = 0.2, COVφ = 0.2
70
70 4 c. COVc = 0.5, COVφ = 0.2
60
60

Pf (%)
2
Pf (%)

50
50
40
40 1
30
30
c
20
20
0.5 b
10 a
10 0.2 0.4
0.1
0.3 0
0 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
0.8 1.0 1.2 1.4 1.6 1.8 2.0
Fs
Fs
(b)
Fig. 3. Probability of failure (Pf) versus deterministic mean value of factor of safety
100
2 2
ðF s Þ with a range of COV for strength parameter values using pLEM [22] and FEM
COVFS = COV + COV
su γ approaches without reinforcement.
90

80 8
70 4 Table 1
Input parameters for FEM analyses with c–/ soil slopes.
60
2
Pf (%)

H:V (slope gradient – horizontal to vertical) 0–1.5:1


50
Slope height, H 10 m
40 1 Cohesion, c (mean) 23 kPa
Friction angle, / (mean) 30°
30 Unit weight, c 20 kN/m3
20 Young’s modulus, E 100 MPa
0.5 Poisson’s ratio, m 0.3
10 0.4
0.1 0.2 0.3
0
0.8 1.0 1.2 1.4 1.6 1.8 2.0 in Table 1. Soil dilation angle was taken as zero in the simulations
Fs to follow based on recommendations by Griffiths and Lane [20].
Fig. 3 shows that both programs gave very similar results for differ-
Fig. 1. Probability of failure (Pf) versus deterministic mean factor of safety ðF s Þ for
ent combinations of COVc and COV/.
cohesive soil slopes with lognormal distribution of undrained shear strength (Su)
and unit weight (c) computed using pLEM with closed-form solution (dashed lines)
and FEM (symbols) approaches for the case of no reinforcement: (a) variability in 3. Modification and verification of mFEM code (reinforced
soil strength (Su) only, (b) variability in soil strength (Su) and soil unit weight (c).
slopes)

3.1. Modification
2H 2H 2H

To the best knowledge of the writers, Matsui and San [31] were
H the first to model a reinforced soil slope using bar elements in a
FEM program that also employed the strength reduction method
2H
applied to the soil strength. In the current investigation three-
H = 10 m node linear elastic bar elements were added to the FEM source
Random variables: Su , γ code (mrslope2d) to model horizontal reinforcement layers. The
bar elements have only one degree of freedom. There is no inter-
face between the reinforcement and soil (i.e. the soil and reinforce-
Fig. 2. Geometry for example cohesive soil slope. ment elements are perfectly bonded). This approach was adopted
for simplicity. Nevertheless, the assumption of a perfect bond is
reasonable for the case of a geogrid product reinforcing a granular
The SVSlope software package [16] was used to carry out circular soil. During pullout the very efficient interlock between the soil
slip analyses (Bishop’s Simplified Method) together with the Float- geogrid apertures forces the failure mechanism to develop in the
ing Method option for probabilistic analyses. The dashed lines in soil adjacent to the geogrid. Yang et al. [42] reported the results
Fig. 3 show numerical results for lc/(lcHtanl/) = 0.2 and l/ = 30°, of FEM modeling of reinforced soil slopes taken to failure and con-
where lc, l/ and lc are mean values of soil cohesion, friction angle cluded that there was no difference in numerical outcomes with
and unit weight. These curves are general and apply also to different and without interface slip elements. FEM simulations reported by
combinations of lc, lc and H, as long as lc/(lcHtanl/) = 0.2. In their Karpurapu and Bathurst [25] satisfactorily reproduced the results
analyses, 4500 Monte Carlo simulations were used in each case. of full-scale geogrid reinforced soil walls taken to failure by uni-
To check the FEM approach for unreinforced cohesive–frictional form surcharging. Their numerical model assumed a perfect bond
soil slopes, a series of analyses were carried out based on the slope between the geogrid and a granular backfill soil. In the current
with the geometry shown in Fig. 2. However, the slope gradient study using the shear strength reduction method, the axial stiff-
(H:V) was varied from 0 to 1.5 to obtain different mean values of ness of the reinforcement is held constant and only the soil
factor of safety. Parameters used in these analyses are summarized strength parameters are reduced.
48 N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55

3.2. Reinforced slope case range of reinforcement stiffness values from J = 100 to 1000 kN/m
and soil friction angles of / = 20° and 40°. Differences in computed
In this investigation a slope with a face angle of 45°, height of Fs values were negligible. These results are expected from deter-
5 m and a horizontal backfill surface and fore slope was assumed ministic stability (ultimate failure limit state) analyses because it
to compare numerical results using different analysis methods. is the strength of the component materials that determine the ulti-
The unfactored friction angle of the soil was assumed as 30°, and mate limit state and not how the failure state was approached.
the unit weight of the soil as 18 kN/m3. To avoid numerical insta- Griffiths and Lane [20] have also shown that the choice of Young’s
bility during finite element analyses and to minimize computation Modulus (E) and Poisson’s ratio (m) has little influence on stability
time, the soil was assigned a small cohesion value of 1 kPa. The analysis outcomes for similar simple conventional slopes. An
possible effect of this small cohesion on numerical outcomes was important implication of these single realization results using the
investigated and shown to be negligible later in the paper. mFEM program in this study is that introducing variability in soil
The factored reinforcement strength, length and spacing used in E, v and reinforcement stiffness will not change probabilistic anal-
the reference soil slope in Fig. 4 were selected using the determin- ysis outcomes for the collapse limit state of slopes that are
istic LEM-based design charts by Bathurst and Jones [8]. The candi- described later in the paper. This will not be the case for probabilis-
date reinforcement was taken as a uniaxial geogrid with properties tic analyses of deformation-based (serviceability) limit states of
reported by Walters et al. [40] that are a typical match to the unreinforced and reinforced slopes.
required factored tensile strength using the design charts of
Bathurst and Jones [8]. The axial stiffness of the reinforcement is
600 kN/m and the factored tensile strength at rupture is 72 kN/m. 3.4. Convergence criterion limit for mFEM code and numerical stability
The latter is the available long-term tensile strength that considers
the reduction in reference tensile strength of the product due to A convergence criterion is implemented in the original FEM and
factors such as installation damage and creep according to conven- modified FEM code developed by the writers to determine when
tional design practice. In conventional 2D FEM programs J = EA analyses are in equilibrium (converged). A solution is deemed to
where E is elastic modulus and A is the bar area per unit width. have converged if the relative change between the largest nodal
Assuming a factor of safety equal to 1.3, the design friction angle displacements of two successive iterations is smaller than a spec-
of soil was 23.9° and the minimum ratio of reinforcement length ified tolerance. Fig. 5 shows that as the convergence tolerance
to slope height was found to be L/H = 0.81, which leads to a mini- decreases, computed reinforcement strains converge to a unique
mum reinforcement length of 4.05 m. For convenience, the length solution. A convergence tolerance of 0.01% was judged to be a suit-
of the reinforcement was taken as 5 m. A similar reinforcement able compromise between accuracy and computational effort and
length was determined using the design charts by Jewell [24]. It was used in all simulation runs using the FEM and mFEM codes.
was determined that a minimum of 5 layers of reinforcement at A similar investigation of sensitivity of reinforcement strain results
1 m vertical spacing was required to achieve a factor of safety of to selected convergence criteria was also carried out using the FEM
1.3 using the charts and spacing recommendations by Bathurst program Sigma/W (described in the next section) to ensure that
and Jones [8]. The geometry of the reinforced slope, reinforcement computed strain values using this program were also unique and
arrangement and matching FEM mesh are shown in Fig. 4. The thus valid comparisons between programs were possible.
reinforced slope design included checks against tensile over- The influence of c = 0 and c = 1 kPa on numerical outcomes was
stressing of the reinforcement layers (i.e. tensile forces greater than examined in a series of simulations using the slope in Fig. 4
72 kN/m) and pullout assuming that the pullout interface shear together with constant friction angle / = 40°. The results showed
strength was equal to that of the soil. These limit states were also that with c = 0 the crest slope displacements were three times
checked independently during LEM and mFEM analyses reported greater at collapse and the simulation took almost three times
later in this study and shown not to be exceeded. longer to reach the same convergence criterion as the case with
c = 1. However, the strength reduction factor at collapse was the
same in both cases. Hence, the small cohesion introduced in this
3.3. Influence of magnitude of soil elastic properties and reinforcement investigation was judged not to change estimates of factor of
stiffness safety. An alternative strategy that was shown to give the same

The same soil modulus (E = 100 MPa) and Poisson’s ratio


(v = 0.3) reported earlier were used in the c–/ soil analyses 0.7
described here. As a check, numerical analyses using the reinforced Convergence tolerance

soil slope in Fig. 4 but with E ranging from 50 to 200 MPa and 0.6 1%
0.1%
Poisson’s ratio ranging from 0.2 to 0.45 were carried out. Differ-
Reinforcement strain (%)

0.01%
ences in numerical outcomes were negligible as expected. A similar 0.5 0.001%
0.0001%
check was carried out using deterministic mFEM analyses with a
0.4

7m 5m 5m 0.3

1 Reinforcement layers
Slope angle: 45°
0.2
2
c = 1 kPa
3 = 30 5m
γ = 18 kN/m3 0.1
7.5 m 4 J = 600 kN/m
1m
5
L=5m 0.0
0 1 2 3 4 5 6
2.5 m
Distance from slope face (m)

Fig. 5. Influence of relative displacement convergence tolerance on magnitude and


Fig. 4. Reinforced slope geometry, reinforcement arrangement and FEM mesh for distribution of reinforcement strains in layer 3 using mFEM code (strength
reinforced slope model. reduction factor = 1.0).
N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55 49

outcome as the case with c = 1 was to set c = 0 and assign a dilation 0.6
angle equal to the friction angle of the soil (associated flow rule). mFEM
0.5 Sigma/W
It may be expected that unlike a purely frictional soil slope, the 0.4
magnitude of soil unit weight will influence numerical outcomes
0.3
when the soil strength includes a cohesive component. A series
0.2
of simulations were carried out for the base case described earlier
Layer 4
and c = 14, 18 and 22 kN/m3 and c = 1 kPa. There were no differ-
0.1

ences in strength reduction factor at collapse using this range of 0.0


0.6
soil unit weights in combination with the very small value of
0.5
c = 1 kPa. Thus numerical outcomes representing collapse were

Strain (%)
0.4
independent of soil unit weight which is the expectation for a
0.3
purely frictional soil.
0.2

3.5. Comparison of program mFEM and Sigma/W strain results from 0.1 Layer 3
deterministic analysis of reinforced soil slope 0.0
0.6
0.5
In order to have confidence that the bar element code added to
the source program was correct, numerical outcomes using the 0.4
mFEM code were compared to results from the well-known FEM 0.3
program Sigma/W [18]. In both cases homogeneous soil properties 0.2
were assumed. Fig. 6 shows that the strains in reinforcement layers 0.1 Layer 2
computed using both programs at end of construction are in good 0.0
agreement. 0 1 2 3 4 5 6
The peak reinforcement strains plotted in Fig. 6 are less than 1%. Distance from slope face (m)
These low values are consistent with measured strains recorded for
Fig. 6. Reinforcement strains in reinforcement layers 2, 3 and 4 using mFEM code
a full-scale reinforced soil embankment at end of construction that and Sigma/W program (strength reduction factor = 1.0).
was built with a similar extensible geogrid reinforcement product
as that assumed in the current investigation and a granular soil fill
0.5
[5] and an instrumented reinforced soil slope wall reported by
Normalized maximum nodal displacement

Maximum number of iterations = 1000


Fannin and Hermann [14]. End-of-construction strains in a large
database of instrumented geosynthetic reinforced soil walls
reported by Allen et al. [2] and Bathurst et al. [4] are most often 1.0
less than 1%. These earlier studies give confidence that the
end-of-construction condition for the reinforcement layers in our
numerical models were reasonable. 1.5

3.6. Determination of factor of safety

2.0
In the shear strength reduction method, a series of increasing Fs = 0.81 (LEM)
user-specified strength reduction factors (e.g. 1.0, 1.01, 1.02. . .)
are used to factor the soil strength parameters down in each FEM Fs = 0.79 (FEM)
analysis cycle. The first strength reduction factor that brings the 2.5
slope to failure is taken as the factor of safety. Slope failure is iden- 0.5 0.6 0.7 0.8 0.9

tified when the finite element model is unable to come to numer- Strength reduction factor
ical convergence within a specified number of iterations. Fig. 7
Fig. 7. Normalized maximum nodal displacement versus strength reduction factor
shows the maximum nodal displacement computed during each (unreinforced slope).
strength reduction computation cycle normalized to the maximum
nodal displacement computed at the first strength reduction trial 1
Normalized maximum nodal displacement

(i.e. the smallest strength reduction factor applied). The results in 2


Maximum number of iterations = 4000
this figure are for an unreinforced purely frictional soil slope with
3
slope geometry shown in Fig. 4. The large change in normalized
maximum nodal displacement occurs at a strength reduction fac- 4
tor of 0.79 which is taken as the minimum (critical) factor of safety 5
for the slope. This value is in reasonable agreement with the factor 6
of safety of 0.81 for the same unreinforced slope using the Bishop’s
7
Simplified Method. Similar agreement was demonstrated for an
example homogeneous c–/ slope by Griffiths and Lane [20] using 8
an earlier version of the same FEM program and the Bishop and 9
Morgenstern Method. They also pointed out that a sufficient 10 Fs = 1.45 (LEM) Fs = 1.50
number of iterations must be selected to ensure that the point of (mFEM)
11
non-convergence is detected and thus an accurate estimate of
the critical strength reduction factor (factor of safety) identified. 12
1.0 1.1 1.2 1.3 1.4 1.5 1.6
For the unreinforced slope cases in the current study, it was deter-
Strength reduction factor
mined that 1000 iterations are sufficient to identify numerical non-
convergence. Similar data are plotted in Fig. 8 for the matching Fig. 8. Normalized maximum nodal displacement versus strength reduction factor
reinforced slope. Here, the maximum nodal displacements increase (reinforced slope).
50 N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55

more smoothly with increasing strength reduction factor than for is expected since the much stiffer soil fails before the reinforce-
the unreinforced slope case. However, the trend of increasing rate ment can strain to rupture. This result is due to the large difference
of maximum nodal displacement can still be identified. It can be in stiffness of typical granular soils and extensible soil reinforce-
seen that failure occurs when the strength reduction factor is equal ment products [1].
to 1.50 and this value is taken as the factor of safety of the rein- In addition to the observations made above, the plots in Figs. 9
forced slope. In this case, the iteration ceiling was set to 4000. and 10 illustrate again that load (or strain) mobilization at end of
The relatively smooth response curve in Fig. 8 (compared to construction is very low. Mobilization of the reinforcement forces
Fig. 7, unreinforced case) is consistent with the expectation that in this simulation initiated at about SRF = 1.08 which matches
reinforced slopes constructed with extensible reinforcement (i.e. the factor of safety of the corresponding unreinforced soil slope.
geosynthetic products) will deform more gradually than unrein- The non-uniform distribution of peak loads in Fig. 10 highlights
forced slopes as failure approaches. This behavior is consistent a fundamental shortcoming of strength-based LEM slope stability
with the observation that the magnitude of the normalized analyses which assume that the tensile force is the same in all lay-
maximum nodal displacement for the reinforced slope at failure ers of a slope constructed with the same reinforcement material.
is larger than for the unreinforced slope case. The data in Fig. 10 show that mobilized forces are greater for the
middle three reinforcement layers. A similar non-uniform
3.7. Mobilization of reinforcement forces using mFEM concave-out distribution of reinforcement forces was observed in
an instrumented geosynthetic reinforced soil slope by Fannin and
Fig. 9 shows the distribution of mobilized tensile force in the Hermann [14].
top reinforcement layer of an example mFEM model slope.
Fig. 10 shows the development of peak mobilized tensile forces 3.8. Comparison of deterministic factor of safety values and failure
for all layers. The plots show that the slope collapsed before the modes for reinforced slopes using mFEM and LEM (Bishop’s Simplified
available strength of any reinforcement layer was mobilized. This Method)

There are two common approaches for the treatment of rein-


8 forcement forces in deterministic limit equilibrium method
Percentage mobilized tensile strength (%)

11
Layer 5 SRF = 1.1
10 (LEM) analysis of reinforced soil slopes. The difference is whether
7 SRF = 1.4
Mobilized tensile force (kN/m)

SRF = 1.8 9 or not the same strength reduction technique applied to soil
SRF = 2.03
6 strength is also applied to the available reinforcement tensile
8
strength [12]. Recall that in the mFEM code used in this study, only
5 7
soil strength is reduced while the stiffness of the reinforcement is
6
4 held constant. Keeping the tensile reinforcement constant is rec-
5 ommended by Duncan et al. [12] for conventional LEM-based slope
3 4 stability analyses. Before comparisons between probabilistic esti-
2 3 mates of failure using the mFEM and pLEM approaches were
2 attempted in the current study, the influence of choice of strength
1
1
reduction method was investigated.
The mFEM and LEM (Bishop’s Simplified Method) were used to
0 0
0 1 2 3 4 5 6 calculate deterministic factors of safety of the reference reinforced
Distance from slope face (m) slope geometry (Fig. 4) using a range of soil friction angles (20–50°).
The factored tensile strength of the reinforcement (72 kN/m)
Fig. 9. Distribution of mobilized tensile force in reinforcement Layer 5 as strength identified earlier was used as the horizontal stabilizing force in all
reduction factor (SRF) is increased to slope collapse (slope angle = 45°, c = 1 kPa,
reinforcement layers in the LEM analyses. Fig. 11 shows that the
/ = 40°, c = 18 kN/m3 and J = 600 kN/m).
mFEM and LEM results for the estimate of the deterministic factor
of safety agree well up to 35° but diverge at larger friction angles.
Percentage mobilized tensile strength (%) Program Slope/W [17] was used to carry out the LEM analyses. In
0 2 4 6 8 10 12 14 16 18 the mFEM and LEM analyses with mobilized reinforcement
5 strength the failure mechanisms move from external type to inter-
nal type as the friction angle increases. An external mechanism is
defined in this study as a failure surface that extends just beyond
4 the reinforced soil zone and an internal mechanism is defined as
one that falls fully within the reinforced soil zone [27]. For the
Height above toe (m)

SRF = 1.10 SRF = 1.40 SRF = 1.80 SRF = 2.03


(Failure) LEM analysis with fixed tensile strength, the failure mechanisms
3
were external type only for all friction angles. Stated alternatively,
the very much higher tensile strength (or force) used in the LEM
2 SRF = 1.00 analyses drives the critical mechanism beyond the reinforcement
SRF = 1.08 layers for all soil friction angle cases. The discrepancy in computed
factor of safety results in Fig. 11 was satisfactorily removed by using
1 the maximum mobilized forces from the mFEM program (Fig. 10) in
the LEM analyses as recommended by Hammah et al. [21].
Fig. 12 shows an example of the internal failure mechanisms
0 predicted from deterministic analysis using mFEM and LEM pro-
0 2 4 6 8 10 12 14
grams with the maximum mobilized reinforcement forces from
Mobilized tensile force (kN/m) the mFEM program used in the LEM analysis. The results are
Fig. 10. Distribution of peak mobilized tensile force in all reinforcement layers as
judged to show good agreement. Fig. 13 shows an example of
strength reduction factor (SRF) is increased to slope collapse (slope angle = 45°, external failure mechanisms using mFEM and LEM programs.
c = 1 kPa, / = 40°, c = 18 kN/m3 and J = 600 kN/m). Again the maximum mobilized reinforcement forces generated
N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55 51

3.5 (a)
mFEM Fs = 1.45 (LEM)
LEM (mobilized tensile strength) Fs = 1.50 (FEM)
3.0 LEM (fixed tensile strength = 72 kN/m)

External failure mechanism only for all φ


2.5 (LEM with fixed tensile strength equal to 72 kN/m)
Fs

2.0
External failure zone Internal failure zone
(mFEM and LEM with (mFEM and LEM with
mobilized tensile strength) mobilized tensile strength)
1.5

1.0
(b)

0.5
15 20 25 30 35 40 45 50 55
Friction angle (degrees)

Fig. 11. Deterministic factor of safety values calculated using LEM (Bishop’s
Simplified Method) and mFEM with progressive mobilized reinforcement tensile
forces and fixed forces equal to the ultimate reinforcement tensile strength of the
reinforcement (LEM – Bishop’s Simplified Method).

(a) (c)
7m 5m 5m
Fs = 2.07 (LEM)
Fs = 2.03 (FEM)
Fs = 1.45 5m

7.5 m

2.5 m

Fig. 13. Example failure mechanisms (external) and factors of safety using mFEM
(Fs = 1.5) (a) plastic displacement vectors, (b) deformed mesh (exaggerated scale),
(b) and (c) LEM (Slope/W with Bishop’s Simplified Method) (Fs = 1.45). Note: c = 1 kPa,
/ = 30°, c = 18 kN/m3 and J = 600 kN/m.

because each Monte Carlo realization in the mFEM analyses will


generate a different set of mobilized maximum forces in the rein-
forcement layers. Hence, the choice of which set of maximum
forces to use in the corresponding pLEM analysis is problematic.

4. Probabilistic stability study of reinforced frictional soil slopes


Using modified FEM code (mFEM)

(c) 7m 5m 5m 4.1. Selection of COV values for / and c, and cross-correlation


coefficient q

Fs = 2.07 5m Based on a review of the literature, Javankhoshdel and Bathurst


[22] identified COVc = 0.1, and COV/ = 0.2 as reasonable upper
7.5 m bounds on the variability in natural soil unit weight and soil
friction angle, respectively. Specific data for granular fills used in
2.5 m constructed slopes and embankments is sparse but may be
expected to be lower. However, to identify trends in numerical
outcomes, COV/ in the range from 0 to 1.0 was also used in some
Fig. 12. Example failure mechanisms (internal) and factors of safety using mFEM sets of analyses in the current study. A positive correlation
(Fs = 2.03) (a) plastic displacement vectors, (b) deformed mesh (exaggerated scale), between / and c is reasonable corresponding to a cross-
and (c) LEM (Slope/W with Bishop’s Simplified Method) (Fs = 2.07). Note: c = 1 kPa,
/ = 40°, c = 18 kN/m3 and J = 600 kN/m.
correlation coefficient in the range of q = 0.1–0.7 [32,10,36,41].
However, to explore possible trends in numerical outcomes, anal-
yses were carried out for values of q = 0.7, 0 and 0.7 in the current
by the mFEM program were used as input in the LEM analysis. The study. The same range of values was also used in probabilistic LEM
results are in good agreement as well. analyses of simple unreinforced soil slopes by Javankhoshdel and
Agreement between mFEM and LEM analyses can be largely Bathurst [23]. Finally, in the analyses to follow a constant value
achieved for deterministic analyses using the method described of c = 1 kPa was assumed; hence the examples are for essentially
above. However, for probabilistic analyses this is not the case purely frictional soils. Recall that this cohesion value was selected
52 N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55

to ensure numerical stability at the free boundaries of the FEM (a)


3.0
domain. With this low value the influence of possible spread in c
values and possible cross-correlations values between random reinforced
values of c and / are of no practical concern. 2.5

4.2. Number of realizations and design target probability of failure 2.0


1.79

The data in the plots to follow were generated using 1000 and unreinforced

Fs
1.5
2000 Monte Carlo simulations for one and two random variable
cases, respectively. For large probabilities of failure (say greater
1.0 0.92
than 1%) these numbers of Monte Carlo simulations are sufficient.
However, large relative differences in Pf < 1% can be expected for
simulations using these numbers and the same case with very 0.5
much larger number of realizations. However, these differences
would not be visually detectable in the figures to follow and there- 0.0
fore are not important if the objective is to illustrate trends in Pf 15 20 25 30 35 40 45 50 55
versus the selected independent variable. Friction angle (degrees)
For actual design of a reinforced slope or embankment the (b)
100
choice of target probability of failure (Pf) is important. A reasonable
value for reinforced soil slopes is 1%. This is the same value recom- 90 COVφ = 0.5, unreinforced
COVφ = 0.5, reinforced
mended for load and resistance factor design (LRFD) calibration of 80 COVφ = 0.2, unreinforced
internal ultimate limit states for reinforced soil walls [3,6,7,28]. COVφ = 0.2, reinforced
70 69
The rationale for this high value is that these systems are highly
66
strength-redundant; if one layer of reinforcement fails, the rein- 60
forcement force can be shed to adjacent reinforcement layers. Pf (%)
50
Since the focus of the paper is on reinforced soil slopes and
embankments, increasing the number of Monte Carlo simulations 40

beyond the values noted was not a practical need. 30


It should be noted that for the design of constructed unrein- 19
20
forced soil slopes and embankments, a lower design probability
of failure Pf = 0.01% has been recommended by Silva et al. [35]. 10
0.6
For these structures and this level of target maximum probability 0
of failure, many more Monte Carlo realizations will be required. 15 20 25 30 35 40 45 50 55
The calculation of the number of realizations to achieve such a Friction angle (degrees)
low Pf value with a satisfactory level of confidence is explained
Fig. 14. (a) Factor of safety (Fs) from deterministic analyses, and (b) probability of
in the textbook by Fenton and Griffiths [15] and is not repeated failure (Pf) versus mean friction angle for unreinforced and reinforced soil slopes
here for brevity. However, some preliminary calculations by the and different COV for soil friction angle (COVc = 0).
writers showed that the required number of realizations using
the FEM code in this paper would be time prohibitive. An alterna- soil friction angle and random variability of soil friction angle on
tive but less accurate strategy is to carry out (say) 1000 Monte probability of failure, respectively. The plots show that for unrein-
Carlo simulations and fit (say) a lognormal function to the distribu- forced and reinforced slopes with the same constant soil properties
tion of computed Fs values [15]. The tail of the fitted distribution and geometry, the reinforced slope has a higher factor of safety and
can then be used to estimate the probability of failure as Pf = p lower probability of failure. For example, for the case of an unrein-
(Fs < 1). The geometry of the slope and/or the soil properties are forced slope with a friction angle of 35°, the factor of safety is 0.92,
then adjusted until computed Pf < 0.01%. which means the slope has failed. On the other hand, a reinforced
An argument in favor of using lower target probabilities of fail- slope with the same soil properties has a factor of safety of 1.79
ure for both unreinforced and reinforced slopes than those noted and is thus stable.
above is the scenario where a long slope or embankment exists. Results of probabilistic analyses are shown in Fig. 14b. For an
The probability of failure of one slope section along a chain of nom- unreinforced slope with a friction angle of 35° and a coefficient
inal identical slopes will increase with the length of the entire of variation COV/ = 0.2, the probability of failure is 69%. If the slope
slope or embankment project. This value can be computed using is constructed with five layers of reinforcement, the probability of
conventional Bernoulli theory knowing the length of the project, failure decreases to 0.6% in this example, which is a very large
a reasonable assumption of the width of a failed slope and the improvement. For the same unreinforced slope but with a coeffi-
probability of failure of the single slope in isolation. cient of variation of 0.5, the probability of failure is 66%. If the slope
is constructed with the same number of reinforcement layers as
4.3. Single random variable, / before, the probability of failure will decrease to 19%. Again, the
reinforcement reduces the probability of failure by a large amount.
In this section, the influence of random variability of soil friction However, the effect of reinforcement on reducing Pf values is also
angle on probabilistic outcomes for unreinforced and reinforced influenced by amount of variability in the estimate of soil friction
slopes is examined. The reinforced purely frictional soil slope angle. For slopes with a low coefficient of variation of /
model shown in Fig. 4 was used here. Reinforcement properties (COV/ = 0.2), the difference between probability of failure values
and slope geometry were held constant. For the matching unrein- of unreinforced and reinforced slopes is relatively large, especially
forced slope, the same slope geometry and soil properties were for friction angles in the range of 25–40°. In particular, for slopes
adopted. with friction angle of 35° or higher the probability of failure is
Fig. 14a and b show results of deterministic analyses using the reduced to practically negligible values by introducing reinforce-
modified FEM code (mFEM) to investigate the influence of mean ment (i.e. less than 0.6%), but this low level of probability of failure
N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55 53

is not possible for reinforced slopes when COV/ = 0.5. These plots 100
demonstrate that for the simple slope cases investigated here, 90
the reinforcement becomes less effective in reducing the probabil- unreinforced
80
ity of failure of a slope as variability in soil properties increases
while mean properties remain unchanged. 70
Fig. 15 shows that for reinforced slopes with four different
60
mean friction angles (30°, 35°, 40°, 45°), the probability of failure

Pf (%)
reinforced
increases as the coefficient of variation of friction angle increases. 50

However, for unreinforced slopes with the same soil properties, 40


the curves exhibit different trends. For unreinforced slopes with
30
friction angles of 40° and 45°, the probability of failure increases
as COV/ increases, which is the same trend as the reinforced 20
COVγ = 0
slopes. However, for unreinforced slope with / = 30° and 35°, the 10 COVγ = 0.1
probability of failure first decreases and then slowly increases with
0
increasing COV/. This behavior can be explained by observing that 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8
both unreinforced slopes with soil friction angle equal to 30° and COVφ
35° have factor of safety values less than 1, meaning that the input
mean values of friction angle are too low for the slopes to be stable Fig. 16. Influence of coefficient of variation of c (COVc) and coefficient of variation
in both cases. When the soil friction angle is treated as a random of friction angle (COV/) on the probability of failure for unreinforced and reinforced
slopes (/ = 30° and q = 0).
variable with small variability (i.e. low COV/), there is only a small
possibility that a large enough value will be randomly selected
during the Monte Carlo simulation process to give an outcome
70
with factor safety greater than 1. For reinforced slopes, all cases Cross-correlation coefficient
have Fs values much greater than 1, hence the probability of failure 60 ρ=0
by selecting friction angles that are small enough to cause failure ρ = 0.7
with the same low variability in friction angle (COV/) is much less. 50 ρ = – 0.7

4.4. Influence of correlated random variables / and c 40


Pf (%)

In this section, both soil friction angle and unit weight are con- 30
sidered as random variables in probabilistic analyses with different COVφ = 0.5
combinations of COV/ and COVc. The effect of cross-correlation 20
between soil friction angle and unit weight on the probabilistic COVφ = 0.2
analysis outcomes of both unreinforced and reinforced slopes is 10
COVφ = 0.1
also investigated.
Fig. 16 shows that the coefficient of variation of unit weight has 0
0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8
no practical influence on the probabilistic outcomes for unrein-
Fs
forced slopes for a wide range of COV/ and uncorrelated / and c.
The small visually detectable difference that is observable in the Fig. 17. Influence of cross-correlation (q) between / and c on the probability of
plots for reinforced and unreinforced soil slopes is due to the small failure for reinforced slope (COVc = 0.1).
cohesion value that was added to the soil in order to ensure
numerical stability. As the value of the cohesive component of soil
The plots in Fig. 17 show that the cross-correlation coefficient
strength goes to zero the influence of magnitude COV of unit
(q) between / and c has almost no influence on the probability
weight will disappear entirely which is expected for reinforced
of failure of the reinforced slope. Again, this is because for
and unreinforced frictional soil slopes.
essentially frictional soil slopes the collapse factor of safety is
independent of soil unit weight and thus independent of any
100
cross-correlation between / and c. The same will be true for
90 unreinforced soil slopes.
φ Fs
80 30
o
0.79
35
o
o
0.92 5. Conclusions
70 40
o
1.08
45 1.26
60 Probabilistic analysis of reinforced soil slopes using limit equi-
Pf (%)

50 30
o
1.50 librium methods (LEMs) and random soil property values have
o
35 1.79
40
o
2.03 been reported in the literature. However, probabilistic analyses
40 45
o
2.28
using the finite element method (FEM) in combination with the
30 strength reduction method have not. This paper is the first attempt
20 and a preliminary effort in this direction to the best knowledge of
the writers.
10 dashed lines: unreinforced cases
solid lines: reinforced cases An open-source FEM program described by Griffiths and Fenton
0 [19] and Fenton and Griffiths [15] was expanded to allow for prob-
0.0 0.2 0.4 0.6 0.8 1.0 1.2
abilistic analysis of reinforced slopes by adding bar elements to
COVφ
model the reinforcement. The slope geometry, reinforcement
Fig. 15. Probability of failure (Pf) versus coefficient of variation of friction angle
arrangement, soil and reinforcement constitutive models have
(COV/) for unreinforced and reinforced slopes with four soil friction angles (/ = 30°, been purposely kept simple in order to focus on the influence of
35°, 40°, 45°) and COVc = 0. random soil variability on factor of safety.
54 N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55

Numerical predictions using the program for both deterministic China Scholarship Council (CSC). Finally, the authors wish to thank
and probabilistic analyses were checked against results from the reviewers whose comments allowed them to clarify a number
closed-form solutions for unreinforced soil, unreinforced and rein- of points in the original submission.
forced LEM-based slope stability programs and another commer-
cially available FEM program.
References
The paper shows that numerical outcomes for deterministic fac-
tor of safety estimates for reinforced slopes using LEM analyses [1] Allen TM, Bathurst RJ. Comparison of working stress and limit equilibrium
may not be the same as those using mFEM analyses. In order to behavior of reinforced soil walls. In: Sound geotechnical research to practice
achieve similar factor of safety outcomes for both internal and Honoring Robert D, Holtz II. ASCE Geotechnical Special Publication No. 230;
2013. p. 500–14.
external failure mechanisms using both approaches for the same [2] Allen TM, Bathurst RJ, Walters DL, Holtz RD, Lee WF. A new working stress
slope, the mobilized reinforcement forces computed from the method for prediction of reinforcement loads in geosynthetic walls. Can
FEM analyses must be used in the LEM analyses. However, this Geotech J 2003;40(5):976–94.
[3] Allen TM, Nowak AS, Bathurst RJ. Calibration to determine load and resistance
general approach does not work for probabilistic analyses using
factors for geotechnical and structural design. Transportation Research Board
mFEM and probabilistic analyses using pLEM. Circular E-C079, Washington, DC; 2005. p. 93.
The major contribution of the work described in this paper is [4] Bathurst RJ, Allen TM, Walters DL. Reinforcement loads in geosynthetic walls
and the case for a new working stress design method. Geotext Geomembr
the extension of the well-known FEM code by Griffiths and Fenton
2005;23(4):287–322.
[19] and Fenton and Griffiths [15] to carry out probabilistic analy- [5] Bathurst RJ, Blatz JA, Burger MH. Performance of full-scale reinforced
ses of reinforced slopes with simple geometry and soil conditions. embankments loaded to failure. Can Geotech J 2003;40(6):1067–83.
The paper demonstrates the utility of the modified FEM program [6] Bathurst RJ, Huang B, Allen TM. Load and resistance factor design (LRFD)
calibration for steel grid reinforced soil walls. Georisk 2011;5(3–4):218–28.
by showing the influence of random soil properties on estimates [7] Bathurst RJ, Huang B, Allen TM. LRFD calibration of the ultimate pullout limit
of probability of failure of reinforced soil slopes. Analysis results state for geogrid reinforced soil retaining walls. ASCE Int J Geomech, Special
demonstrate the large reduction in probability of failure that can Issue Geosynth 2012;12(4):399–413.
[8] Bathurst RJ, Jones CJFP. Earth retaining structures and reinforced slopes. In:
be realized by adding (geosynthetic) reinforcement layers to con- Rowe RK, editor. Geotechnical and geoenvironmental engineering
structed slopes. Another advantage of the mFEM approach over handbook. Norwell, MA, USA: Kluwer Academic Publishing; 2001. p. 1088
pLEM is that probabilistic analyses of simple slopes of the type [chapter 17].
[9] Cho SE. Probabilistic assessment of slope stability that considers the spatial
examined in the current study are not constrained by failure mech- variability of soil properties. J Geotech Geoenviron Eng 2010;136(7):975–84.
anism type and reinforcement forces that must be assumed a priori [10] Chowdhury RN, Xu DW. Reliability index for slope stability assessment – two
using pLEM. However, the utility of the mFEM and FEM approaches methods compared. Reliab Eng Syst Saf 1992;37(2):99–108.
[11] Duncan JM. Factors of safety and reliability in geotechnical engineering. J
decreases if large numbers of realizations are required to achieve
Geotech Geoenviron Eng 2000;126(4):307–16.
very low probabilities of failure. [12] Duncan JM, Wright SG, Brandon TL. Soil strength and slope stability. 2nd
The results of slope analyses with essentially purely frictional ed. New York, NY, USA: John Wiley & Sons; 2014.
[13] El-Ramly H, Morgenstern NR, Cruden DM. Probabilistic slope stability analysis
soils showed that probability of failure is independent of unit
for practice. Can Geotech J 2002;39(3):665–83.
weight and cross-correlation between unit weight and friction [14] Fannin RJ, Hermann S. Performance data for a sloped reinforced soil wall. Can
angle, which is expected. However, probability of failure was sen- Geotech J 1990;27(5):676–86.
sitive to uncertainty in soil friction angle. Nevertheless, variability [15] Fenton GA, Griffiths DV. Risk assessment in geotechnical engineering. New
York, NY, USA: John Wiley & Sons; 2008.
in friction angle for constructed cohesionless granular fills may be [16] Fredlund MD, Thode R. SVSlope theory manual. Saskatoon, Saskatchewan,
very low (e.g. less than COV/ = 0.1). Canada: SoilVision Systems Inc.; 2011.
While not explored in the current investigation, the same mod- [17] Geo-Slope Ltd.. Slope/W for slope stability analysis: user’s guide. Calgary,
Canada: Geo-Slope International Ltd.; 2012.
ified FEM code can also be used to investigate the possible influ- [18] Geo-Slope Ltd.. Stress-deformation modeling with Sigma/W. Calgary,
ence of soil spatial variability on probability of failure of Canada: Geo-Slope International Ltd.; 2014.
constructed unreinforced and reinforced slopes and embankments [19] Griffiths DV, Fenton GA. Probabilistic slope stability analysis by finite
elements. J Geotech Geoenviron Eng 2004;130(5):507–18.
since this capability has been retained in the updated code. How- [20] Griffiths DV, Lane PA. Slope stability analysis by finite elements. Geotechnique
ever, the scale of fluctuations (at least in the vertical direction) is 1999;49(3):387–403.
expected to be at lift thickness dimensions for constructed fills [21] Hammah R, Yacoub T, Curran J. Investigating the performance of the shear
strength reduction (SSR) method on the analysis of reinforced slopes. In:
(say 150–300 mm). The very small scale of fluctuations together
Proceedings of 59th Canadian geotechnical conference, Vancouver, Canadian
with small COV values of soil properties in constructed granular fill Geotechnical Society; 2006. p. 371–5.
slopes (e.g. COVc and COV/ 6 10%) means that uncertainty in soil [22] Javankhoshdel S, Bathurst RJ. Simplified probabilistic slope stability design
charts for cohesive and c-/ soils. Can Geotech J 2014;51(9):1033–45.
properties may be satisfactorily addressed assuming homogenous
[23] Javankhoshdel S, Bathurst RJ. Influence of cross-correlation between soil
random soil properties for the simple slope geometries in this parameters on probability of failure of simple cohesive and c–/ slopes. Can
study. Recall also that project specifications for constructed fills Geotech J 2016:53. http://dx.doi.org/10.1139/cgj–2015-010.
include tight controls on variability in compacted fills to ensure [24] Jewell RA. Application of revised design charts for steep reinforced slopes.
Geotext Geomembr 1991;10(3):203–33.
that there are not significant random and spatial variations in fill [25] Karpurapu RG, Bathurst RJ. Behaviour of geosynthetic reinforced soil retaining
properties. The level of FEM mesh discretization required to carry walls using the finite element method. Comput Geotech 1995;17(3):279–99.
out simulations with soil property fluctuations at lift thickness [26] Kitch W. Deterministic and probabilistic analyses of reinforced soil slopes. PhD
dissertation, University of Texas, Austin, TX; 1994.
scale is likely prohibitive. Notwithstanding the points made above, [27] Kitch WA, Gilbert RB, Wright SG. Probabilistic assessment of commercial
investigation of the influence of spatial variability for constructed design guides for steep reinforced slopes: implications for design. In: ASCE
unreinforced and reinforced simple slopes is a possible future line proceedings of Georisk 2011, Atlanta, Georgia. Reston, VA; 2011 [d 20110000].
[28] Lazarte CA. Proposed specifications for LRFD soil-nailing design and
of investigation. construction. NCHRP report 701, Transportation Research Board, National
Research Council, Washington, DC; 2011.
[29] Leshchinsky D, Boedeker RH. Geosynthetic reinforced soil structures. J Geotech
Acknowledgments
Eng 1989;115(10):1459–78.
[30] Low BK, Tang WH. Reliability analysis of reinforced embankments on soft
The authors are grateful for financial support through a ground. Can Geotech J 1997;34(5):672–85.
research grant awarded to the second author by the Natural [31] Matsui T, San K. Finite element slope stability analysis by shear strength
reduction technique. Soils Found 1992;32(1):59–70.
Sciences and Engineering Research Council of Canada (NSERC). [32] Matsuo M, Kuroda K. Probabilistic approach to design of embankments. Soils
The first author is also grateful for the financial support from the Found 1974;14(2):1–17.
N. Luo et al. / Computers and Geotechnics 77 (2016) 45–55 55

[33] Phoon KK, Kulhawy FH. Characterization of geotechnical variability. Can [38] Tang WH, Yucemen MS, Ang AH-SC. Probability-based short term design of soil
Geotech J 1999;36(4):612–24. slopes. Can Geotech J 1976;13(3):201–15.
[34] Schneider HR, Holtz RD. Design of slopes reinforced with geotextiles and [39] Taylor DW. Stability of earth slopes. Boston Soc Civil Eng 1937;24(3):197–246.
geogrids. Geotext Geomembr 1986;3(1):29–51. [40] Walters DL, Allen TM, Bathurst RJ. Conversion of geosynthetic strain to load
[35] Silva F, Lambe TW, Marr WA. Probability and risk of slope failure. ASCE J using reinforcement stiffness. Geosynth Int 2002;9(5–6):483–523.
Geotech Geoenviron Eng 2008;134(12):1691–9. [41] Wu XZ. Trivariate analysis of soil ranking-correlated characteristics and its
[36] Sivakumar Babu GL, Srivastava A. Reliability analysis of allowable pressure on application to probabilistic stability assessments in geotechnical engineering
shallow foundation using response surface method. Comput Geotech 2007;34 problems. Soils Found 2013;53(4):540–56.
(3):187–94. [42] Yang K-H, Zornberg JG, Liu C-N, Lin H-D. Stress distribution and development
[37] Smith IM, Griffiths DV, Margetts L. Programming the finite element method. within geosynthetic-reinforced soil slopes. Geosynth Int 2012;19(1):62–78.
5th ed. Chichester, UK: John Wiley & Sons; 2014.

You might also like