You are on page 1of 20

Griffiths, 2004) are combined with random-field generation

techniques. This method, called the random finite-element


method (RFEM), fully accounts for spatial correlation and
averaging and is also a powerful slope stability analysis
tool that does not require a priori assumptions relating to
CHAPTER 13 the shape or location of the failure mechanism.
This chapter applies the RFEM to slope stability risk
assessment. Although the authors have also considered
c slopes (Szynakiewicz et al., 2002), the next section
considers a cohesive soil and investigates the general prob-
Slope Stability abilistic nature of a slope. The final section develops a
risk assessment model for slopes. Both sections employ
the RFEM program called RSLOPE2D to perform the
slope stability simulations. This program is available at
http://www.engmath.dal.ca/rfem.

13.1 INTRODUCTION 13.2 PROBABILISTIC SLOPE STABILITY


Slope stability analysis is a branch of geotechnical engi- ANALYSIS
neering that is highly amenable to probabilistic treatment In order to demonstrate some of the benefits of RFEM and
and has received considerable attention in the literature. to put it in context, this section investigates the probabilis-
The earliest studies appeared in the 1970s (e.g., Matsuo and tic stability characteristics of a cohesive slope using both
Kuroda, 1974; Alonso, 1976; Tang et al., 1976; Vanmarcke, simple and more advanced methods. Initially, the slope is
1977) and have continued steadily (e.g., DAndrea and San- investigated using simple probabilistic concepts and classi-
grey, 1982; Li and Lumb, 1987; Mostyn and Li, 1993; cal slope stability techniques, followed by an investigation
Chowdhury and Tang, 1987; Whitman, 2000; Wolff, 1996; on the role of spatial correlation and local averaging. Fi-
Lacasse, 1994; Christian et al., 1994; Christian, 1996; La- nally, results are presented from a full RFEM approach.
casse and Nadim, 1996; Hassan and Wolff, 2000; Duncan, Where possible throughout this section, the probability of
2000; Szynakiewicz et al., 2002; El-Ramly et al., 2002; failure (pf ) is compared with the traditional factor of safety
Griffiths and Fenton, 2004; Griffiths et al., 2006, 2007).
(FS ) that would be obtained from charts or classical limit
Two main observations can be made in relation to the ex-
equilibrium methods.
isting body of work on this subject. First, the vast majority
The slope under consideration, denoted the test problem,
of probabilistic slope stability analyses, while using novel
is shown in Figure 13.1 and consists of undrained clay, with
and sometimes quite sophisticated probabilistic method-
shear strength parameters u = 0 and cu . In this study, the
ologies, continue to use classical slope stability analysis
slope inclination and dimensions given by , H , and D and
techniques (e.g., Bishop, 1955) that have changed little in
the saturated unit weight of the soil sat are held constant,
decades and were never intended for use with highly vari-
able soil shear strength distributions. An obvious deficiency while the undrained shear strength cu is assumed to be a
of the traditional slope stability approaches is that the shape random variable. In the interests of generality, the undrained
of the failure surface (e.g., circular) is often fixed by the shear strength will be expressed in dimensionless form c,
method; thus the failure mechanism is not allowed to seek where c = cu /(sat H ).
out the most critical path through the soil. Second, while
the importance of spatial correlation (or autocorrelation) 13.2.1 Probabilistic Description of Shear Strength
and local averaging of statistical geotechnical properties has
long been recognized by many investigators (e.g., Mostyn In this study, the shear strength c is assumed to be char-
and Soo, 1992), it is still regularly omitted from many prob- acterized statistically by a lognormal distribution defined
abilistic slope stability analyses. by a mean c and a standard deviation c . Figure 13.2
In recent years, the authors have been pursuing a more shows the distribution of a lognormally distributed cohesion
rigorous method of probabilistic geotechnical analysis (e.g., having mean c = 1 and standard deviation c = 0.5. The
Griffiths and Fenton, 2000a; Paice, 1997), in which nonlin- probability of the strength dropping below a given value
ear finite-element methods (program 6.3 from Smith and can be found from standard tables by first transforming the

Risk Assessment in Geotechnical Engineering Gordon A. Fenton and D. V. Griffiths 381


Copyright 2008 John Wiley & Sons, Inc. ISBN: 978-0-470-17820-1
382 13 SLOPE STABILITY

Input parameters A third parameter, the spatial correlation length ln c , will


2 also be considered in this study. Since the actual undrained
1 fu = 0, gsat
H
shear strength field is lognormally distributed, its logarithm
mcu, scu, qln cu yields an underlying normal distributed (or Gaussian)
D=2 b
DH field. The spatial correlation length is measured with respect
b = 26.6
to this underlying field, that is, with respect to ln c. In
particular, the spatial correlation length (ln c ) describes the
distance over which the spatially random values will tend
to be significantly correlated in the underlying Gaussian
Figure 13.1 Cohesive slope test problem. field. Thus, a large value of ln c will imply a smoothly
varying field, while a small value will imply a ragged
field. The spatial correlation length can be estimated from
1.5

a set of shear strength data taken over some spatial region


simply by performing the statistical analyses on the log-
data. In practice, however, ln c is not much different in
magnitude from the correlation length in real space, and, for
Mode = 7.16 most purposes, c and ln c are interchangeable given their
Median = 8.94 inherent uncertainty in the first place. In the current study,
1.0

Mean = 1.0 the spatial correlation length has been nondimensionalized


by dividing it by the height of the embankment H and will
be expressed in the form
f(c)

ln c
= (13.4)
H
0.5

It has been suggested (see, e.g., Lee et al., 1983; Kul-


hawy et al., 1991) that typical vc values for undrained
shear strength lie in the range 0.10.5. The spatial cor-
relation length, however, is less well documented and may
well exhibit anisotropy, especially since soils are typically
horizontally layered. While the advanced analysis tools
0

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 used later in this study have the capability of modeling
c an anisotropic spatial correlation field, the spatial corre-
lation, when considered, will be assumed to be isotropic.
Figure 13.2 Lognormal distribution with mean 1 and standard Anisotropic site-specific applications are left to the reader.
deviation 0.5 (vc = 0.5).
13.2.2 Preliminary Deterministic Study
lognormal to the normal: To put the probabilistic analyses in context, an initial
  deterministic study has been performed assuming a uniform
ln a ln c soil. By a uniform soil we mean that the soil properties are
P[c < a] = P[ln c < ln a] = P Z <
ln c the same at all points through the soil mass. For the simple
 
ln a ln c slope shown in Figure 13.1, the factor of safety FS can
= (13.1) readily be obtained from Taylors (1937) charts or simple
ln c
limit equilibrium methods to give Table 13.1.
as is discussed in Section 1.10.9. The lognormal parame-
ters ln c and ln c given c and c are obtained via the
transformations Table 13.1 Factors of Safety for Uniform Soil
  c FS
ln2 c = ln 1 + vc2 (13.2a)
0.15 0.88
ln c = ln(c ) 12 ln2 c (13.2b) 0.17 1.00
in which the coefficient of variation of c, vc , is defined as 0.20 1.18
c 0.25 1.47
vc = (13.3) 0.30 1.77
c
PROBABILISTIC SLOPE STABILITY ANALYSIS 383
where  is the cumulative standard normal distribution
3

function (see Section 1.10.8).


This approach has been repeated for a range of c and
2.5

vc values, for the slope under consideration, leading to


Figure 13.4, which gives a direct relationship between the
FS and the probability of failure. It should be emphasized
2

that the FS in this plot is based on the value that would


have been obtained if the slope had consisted of a uniform
soil with a shear strength equal to the mean value c from
1.5
FS

Figure 13.3. We shall refer to this as the factor of safety


FS = 1 based on the mean.
From Figure 13.4, the probability of failure pf clearly
1

increases as the FS decreases; however, it is also shown


that for FS > 1, the probability of failure increases as the vc
c = 0.17
0.5

increases. The exception to this trend occurs when FS < 1.


As shown in Figure 13.4, the probability of failure in such
cases is understandably high; however, the role of vc is to
0

0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 0.5 have the opposite effect, with lower values of vc tending
c to give the highest values of the probability of failure. This
is explained by the bunching up of the shear strength
Figure 13.3 Linear relationship between FS and c for uniform
distribution at low vc rapidly excluding area to the right of
cohesive slope with slope angle = 26.57 and depth ratio D = 2.
the critical value of c = 0.17.
Figure 13.5 shows that the median (see Section 1.6.2),
These results, shown plotted in Figure 13.3, indicate the c is the key to understanding how the probability of fail-
linear relationship between c and FS . The figure also shows ure changes in this analysis. When c < 0.17, increasing
that the test slope becomes unstable when the shear strength vc causes pf to fall, whereas when c > 0.17, increasing vc
parameter falls below c = 0.17. The depth ratio mentioned causes pf to rise.
in Figure 13.3 is defined in Figure 13.1.

13.2.3 Single-Random-Variable Approach


1.1

The first probabilistic analysis to be presented here in-


1

vestigates the influence of giving the shear strength c a


0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

lognormal probability density function similar to that shown


in Figure 13.2, based on a mean c and a standard devi-
ation c . The slope is assumed to be uniform, having the
same value of c everywhere; however, the value of c is
selected randomly from the lognormal distribution. Antic-
ipating the random-field analyses to be described later in
pf

this section, this single-random-variable (SRV) approach


implies a spatial correlation length of  = .
The probability of failure (pf ) in this case is simply equal pf = 0.28 uc = 0
to the probability that the shear strength parameter c will uc = 0.125
uc = 0.25
be less than 0.17. Quantitatively, this equals the area of the uc = 0.5
FS = 1.47

probability density function corresponding to c 0.17. uc = 1


uc = 2
For example, if c = 0.25 and c = 0.125 (vc = 0.5), uc = 4
0.1 0

Eqs. 1.176 state that the mean and standard deviation of the uc = 8
underlying normal distribution of the strength parameter are
0.8 1 1.2 1.4 1.6 1.8 2 2.2
ln c = 1.498 and ln c = 0.472.
FS
The probability of failure is therefore given by
 
ln 0.17 ln c Figure 13.4 Probability of failure versus FS (based on mean) in
pf = p[c < 0.17] =  = 0.281 SRV approach.
ln c
384 13 SLOPE STABILITY

1
1
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

0.9
f1 = 0.0
f1 = 0.2

0.8
f1 = 0.4
f1 = 0.6

0.7
= 0.1
m c
= 0.15

0.6
m c
= 0.17
m
pf

c
= 0.2
m

0.5
pf
c

mc = 0.25

mc = 0.3

0.4
0.3
0.2
0.1
0

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


uc

0
2 3 4 5 6 7 89 2 3 4 5 6 7 89
Figure 13.5 Probability of failure pf versus coefficient of vari- 101 100 101
ation vc for different medians of c, c . uc
(a)
1

While the SRV approach described in this section leads to


simple calculations and useful qualitative comparisons be-
0.9

tween the probability of failure and the FS , the quantitative f2 = 0.0


f2 = 0.125
0.8

value of the approach is more questionable. An important f2 = 0.25


observation highlighted in Figure 13.4 is that a soil with f2 = 0.5
0.7

a mean strength of c = 0.25 (implying FS = 1.47) would f2 = 1.0

give a probability of failure as high as pf = 0.28 for a


0.6

soil with vc = 0.5. Practical experience indicates that slopes


0.5
pf

with an FS as high as 1.47 rarely fail.


An implication of this result is that either the perfectly
0.4

correlated SRV approach is entirely pessimistic in the pre-


0.3

diction of the probability of failure, and/or it is unconserva-


tive to use the mean strength of a variable soil to estimate
0.2

the FS . Presented with a range of shear strengths at a given


site, a geotechnical engineer would likely select a pes-
0.1

simistic or lowest plausible value for design, cdes , that


0

would be lower than the mean. Assuming for the time be- 2 3 4 5 6 7 89 2 3 4 5 6 7 89
ing that the SRV approach is reasonable, Figure 13.6 shows 101 100 101
the influence on the probability of failure of two strategies uc
for factoring the mean strength c prior to calculating the (b)
FS for the test problem. In Figure 13.6a, a linear reduction
Figure 13.6 Influence of different design strength factoring
in the design strength has been proposed using a strength
strategies on probability of failureFS relationship: (a) linear
reduction factor f1 , where
factoring and (b) standard deviation factoring; all curves assume
cdes = c (1 f1 ) (13.5) FS = 1.47 (based on cdes = 0.25).
and in Figure 13.6b, the design strength has been reduced
from the mean by a factor f2 of the standard deviation,
probability of failure of pf = 0.28 with no strength factor-
where
ization, f1 = f2 = 0, has also been highlighted for the case
cdes = c f2 c (13.6)
of vc = 0.5. In both plots, an increase in the strength reduc-
All the results shown in Figure 13.6 assume that after tion factor reduces the probability of failure, which is to be
factorization, cdes = 0.25, implying an FS of 1.47. The expected; however, the nature of the two sets of reduction
PROBABILISTIC SLOPE STABILITY ANALYSIS 385
curves is quite different, especially for higher values of

1
vc . From the linear mean strength reduction (Eq. 13.5),
f1 = 0.6 would result in a probability of failure of about
0.6%. By comparison, a mean strength reduction of one

0.8
standard deviation given by f2 = 1 (Eq. 13.6) would result
in a probability of failure of about 2%. Figure 13.6a shows
a gradual reduction of the probability of failure as f1 is

0.6
increased; however, a quite different behavior is shown in
Figure 13.6b, where standard deviation factoring results in

r(t)
a very rapid reduction in the probability of failure, espe-

0.4
cially for higher values of vc > 2. This curious result is
easily explained by the functional relationship between pf
and vc , where the design strength can be written as

0.2
cdes = 0.25 = c f2 c = c (1 f2 vc ) (13.7) r = 0.135

Hence as vc 1/f2 , c . With the mean strength so


much greater than the critical value of 0.17, the probability

0
of failure falls very rapidly toward zero. 0 0.5 1 1.5 2
t/qln c

13.2.4 Spatial Correlation Figure 13.7 Markov correlation function.

Implicit in the SRV approach described above is that the


spatial correlation length is infinite. In other words only 13.2.5 Random Finite-Element Method
uniform soils are considered in which the single property
assigned to the slope is taken at random from a lognormal A powerful and general method of accounting for spatially
distribution. A more realistic model would properly take random shear strength parameters and spatial correlation
is the RFEM, which combines elasto-plastic finite-element
account of smaller spatial correlation lengths in which the
analysis with random-field theory generated using the LAS
soil strength is allowed to vary spatially within the slope.
method (Section 6.4.6). The methodology has been de-
The parameter that controls this behavior (at least under the
scribed in more detail in previous chapters, so only a brief
simple spatial variability models considered here) is the spa-
description will be repeated here.
tial correlation length ln c as discussed previously. In this
A typical finite-element mesh for the test problem con-
work, an exponentially decaying (Markovian) correlation
sidered in this section is shown in Figure 13.8. The majority
function is used of the form
of the elements are square; however, the elements adjacent
( ) = e 2| |/ln c (13.8) to the slope are degenerated into triangles.
The code developed by the authors enables a random field
where ( ) is the familiar correlation coefficient between of shear strength values to be generated and mapped onto
two points in the soil mass which are separated by dis- the finite-element mesh, taking full account of element size
tance . A plot of this function is given in Figure 13.7 in the local averaging process. In a random field, the value
and indicates, for example, that the soil strength at two assigned to each cell (or finite element in this case) is itself
points separated by = ln c (/ln c = 1) will have a cor- a random variable; thus, the mesh of Figure 13.8, which
relation coefficient of = 0.135. This correlation function
is merely a way of representing the observation that soil
samples taken close together are more likely to have sim- 2H 2H 2H
ilar properties than samples taken from far apart. There is
also the issue of anisotropic spatial correlation in that soils Unit weight g H
are likely to have longer spatial correlation lengths in the
horizontal direction than in the vertical, due to the depo-
H
sitional history. While the tools described in this section
can take account of anisotropy, this refinement is left to the
reader for site-specific refinements. Figure 13.8 Mesh used for RFEM slope stability analysis.
386 13 SLOPE STABILITY

For a given set of input shear strength parameters (mean,


H standard deviation, and spatial correlation length), Monte
Carlo simulations are performed. This means that the slope
stability analysis is repeated many times until the statistics
of the output quantities of interest become stable. Each
realization of the Monte Carlo process differs in the
(a)
locations at which the strong and weak zones are situated.
For example, in one realization, weak soil may be situated
in the locations where a critical failure mechanism develops
H causing the slope to fail, whereas in another, strong soil in
those locations means that the slope remains stable.
In this study, it was determined that 1000 realizations
of the Monte Carlo process for each parametric group was
sufficient to give reliable and reproducible estimates of the
(b)
probability of failure, which was simply defined as the
proportion of the 1000 Monte Carlo slope stability analyses
Figure 13.9 Influence of correlation length in RFEM analysis: that failed.
(a)  = 0.2; (b)  = 2.0. In this study,failure was said to have occurred if, for
any given realization, the algorithm was unable to converge
within 500 iterations. While the choice of 500 as the
has 910 finite elements, contains 910 random variables.
iteration ceiling is subjective, Figure 13.10 confirms, for the
The random variables can be correlated to one another by
case of c = 0.25 and  = 1, that the probability of failure
controlling the spatial correlation length ln c as described
defined this way, is stable for iteration ceilings greater than
previously; hence, the SRV approach discussed in the
about 200.
previous section, where the spatial correlation length is
implicitly set to infinity, can now be viewed as a special
case of a much more powerful analytical tool. Figures 13.9a 13.2.6 Local Averaging
and 13.9b show typical meshes corresponding to different The input parameters relating to the mean, standard devia-
spatial correlation lengths. Figure 13.9a shows a relatively tion, and spatial correlation length of the undrained strength
low spatial correlation length of  = 0.2, and Figure 13.9b
shows a relatively high spatial correlation length of  = 2.
Light regions depict weak soil. It should be emphasized
1

that both these shear strength distributions come from the


0.9

same lognormal distribution, and it is only the spatial


correlation length that is different.
0.8

In brief, the analyses involve the application of grav-


ity loading and the monitoring of stresses at all the Gauss
0.7

points. The slope stability analyses use an elastic-perfectly


0.6

plastic stressstrain law with a Tresca failure criterion


mc = 0.25
which is appropriate for undrained clays. If the Tresca
0.5

=1
pf

criterion is violated, the program attempts to redistribute ex-


uc = 0.125
cess stresses to neighboring elements that still have reserves
0.4

uc = 0.25
of strength. This is an iterative process which continues un- uc = 0.5
uc = 1
0.3

til the Tresca criterion and global equilibrium are satisfied


uc = 2
at all points within the mesh under quite strict tolerances.
0.2

Plastic stress redistribution is accomplished using a vis-


coplastic algorithm with 8-node quadrilateral elements and
0.1

reduced integration in both the stiffness and stress redistri-


0

bution parts of the algorithm. The theoretical basis of the


0 50 100 150 200 250 300 350 400 450 500 550 600
method is described more fully in Chapter 6 of the text by
Iteration ceiling
Smith and Griffiths (2004), and for a detailed discussion of
the method applied to slope stability analysis, the reader is Figure 13.10 Influence of plastic iteration ceiling on computed
referred to Griffiths and Lane (1999). probability of failure.
PROBABILISTIC SLOPE STABILITY ANALYSIS 387
are assumed to be defined at the point level. While statis- assumed the same in any direction for simplicity):
tics at this resolution are obviously impossible to measure  
2
in practice, they represent a fundamental baseline of the in- (1 , 2 ) = exp + 2
2 2
(13.10)
ln c 1
herent soil variability which can be corrected through local
averaging to take account of the sample size. where 1 is the difference between the x1 coordinates of
In the context of the RFEM approach, each element is as- any two points in the random field, and 2 is the difference
signed a constant property at each realization of the Monte between the x2 coordinates. We assume that x1 is measured
Carlo process. The assigned property represents an average in the horizontal direction and x2 is measured in the vertical
over the area of each finite element used to discretize the direction.
For a square finite element of side length ln c as shown
slope. If the point distribution is normal, local arithmetic
in Figure 13.11, so that A = ln c ln c , it can be shown
averaging is used which results in a reduced variance but
(Vanmarcke, 1984) that for an isotropic spatial correlation
the mean is unaffected. In a lognormal distribution, how-
field, the variance reduction factor is given by
ever, local geometric averaging is used (see Section 4.4.2),
and both the mean and the standard deviation are reduced
ln c
ln c
4
by this form of averaging as is appropriate for situations in (A) = (ln c x1 )(ln c x2 )
(ln c )4 0 0
which low-strength regions dominate the effective strength. 
The reduction in both the mean and standard deviation is 2 2
exp x +y 2 dx1 dx2 (13.11)
because from Eqs. 1.175a and 1.175b, the mean of a log- ln c
normally random variable depends on both the mean and Numerical integration of this function leads to the variance
the variance of the underlying normal log-variable. Thus, reduction values given in Table 13.2 and shown plotted in
the coarser the discretization of the slope stability prob- Figure 13.11.
lem and the larger the elements, the greater the influence Figure 13.11 indicates that elements that are small rel-
of local averaging in the form of a reduced mean and stan- ative to the correlation length ( 0) lead to very little
dard deviation. These adjustments to the points statistics are variance reduction [ (A) 1], whereas elements that are
fully accounted for in the RFEM and are implemented be- large relative to the correlation length can lead to very sig-
fore the elasto-plastic finite-element slope stability analysis nificant variance reduction [ (A) 0].
takes place. The statistics of the underlying log-field, including local
arithmetic averaging, are therefore given by

13.2.7 Variance Reduction over Square Finite ln cA = ln c (A) (13.12a)
Element
In this section, the algorithm used to compute the locally
1

averaged statistics applied to the mesh is described. A


lognormal distribution of a random variable c, with point
statistics given by a mean c , a standard deviation c , and
0.8

spatial correlation length ln c is to be mapped onto a mesh


of square finite elements. Each element will be assigned a
0.6

single value of the undrained strength parameter.


The locally averaged statistics over the elements will be aqln c
g(A)

referred to here as the area statistics with the subscript A.


0.4

Thus, with reference to the underlying normal distribution


of ln c, the mean, which is unaffected by local averaging, is aqln c
given by ln cA , and the standard deviation, which is affected
0.2

by local averaging is given by ln cA .


The variance reduction factor due to local averaging
is defined as (see also Section 3.4)
0

  2 4 68 2 4 68 2 4 68 2 4 68 2 4 68
ln cA 2 103 102 101 100 101 102
(A) = (13.9)
ln c a

and is a function of the element size, A, and the correlation Figure 13.11 Variance reduction when arithmetically averaging
function from Eq. 13.8, repeated here explicitly for the two- over square element of side length ln c with Markov correlation
dimensional isotropic case (i.e., the correlation length is function (A = ln c ln c ).
388 13 SLOPE STABILITY

Table 13.2 Variance Reduction due to Arithmetic

1
Averaging over Square Element
(A)

0.8
0.01 0.9896
0.10 0.9021

ucA /uc
0.6
1.00 0.3965
10.00 0.0138 uc = 0.1

0.4
uc = 0.5
uc = 1.0

0.2
and
ln cA = ln c (13.12b)

0
0 2 4 6 8 10
which leads to the following statistics of the lognormal a
field, including local geometric averaging, that is actually (a)
mapped onto the finite-element mesh (from Eqs. 1.175a and

1
1.175b)
0.8
cA = exp ln c + 12 ln2 c (A) (13.13a)

cA = cA exp{ln2 c (A)} 1
0.6

(13.13b)
ucA /uc

from which it is easy to see that local geometric averaging


0.4

uc = 0.1
affects both the mean and the standard deviation. Recall uc = 0.5
also that arithmetic averaging of ln c corresponds to geo- uc = 1.0
0.2

metric averaging of c (see Section 4.4.2 for more details).


It is instructive to consider the range of locally averaged
statistics since this helps to explain the influence of the
0

spatial correlation length (= ln c /H ) on the probability 0 2 4 6 8 10


a
of failure in the RFEM slope analyses described in the next
(b)
section.
Expressing the mean and the coefficient of variation of Figure 13.12 Influence of element size, expressed in the form
the locally averaged variable as a proportion of the point of size parameter , on statistics of local averages: influence on
values of these quantities leads to Figures 13.12a and the (a) mean and (b) coefficient of variation.
13.12b, respectively. In both cases, there is virtually no
reduction due to local averaging for elements that are small
relative to the spatial correlation length ( 0). This is to variable can be written as
be expected since the elements are able to model the point
vcA (1 + vc2 ) (A) 1
field quite accurately. For larger elements relative to the = (13.15)
spatial correlation length, however, Figure 13.12a indicates vc vc
that the average of the locally averaged field tends to a which states that when (A) 0, vcA /vc 0, thus
constant equal to the median, and Figure 13.12b indicates vcA 0.
that the coefficient of variation of the locally averaged field Further examination of Eqs. 13.14 and 13.15 shows that
tends to zero. for all values of (A) the median of the geometric average
From Eqs. 13.12 and 13.13, the expression plotted in equals the median of c:
Figure 13.12a for the mean can be written as cA = c (13.16)
cA 1
= (13.14) Hence it can be concluded that:
c (1 + vc2 )[1 (A)]/2

which states that when (A) 0, cA /c 1/ 1 + Vc2 , 1. Local geometric averaging reduces both the mean and
thus cA e ln c = c , which is the median of c. Simi- the variance of a lognormal point distribution.
larly, the expression plotted in Figure 13.12b for the co- 2. Local geometric averaging preserves the median of
efficient of variation of the locally geometrically averaged the point distribution.
PROBABILISTIC SLOPE STABILITY ANALYSIS 389
3. In the limit as A and/or  0, local geometric

1.1
averaging removes all variance, and the mean tends

1
to the median.

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9


13.2.8 Locally Averaged SRV Approach
In this section the probability of failure is reworked with
the SRV approach using properties derived from local
averaging over an individual finite element, termed finite-

pf
element locally averaged properties throughout the rest = 0.0
of this section. With reference to the mesh shown in = 0.03125
= 0.0625
Figure 13.8, the square elements have a side length of = 0.125
0.1H , thus  = 0.1/. Figure 13.13 shows the probability =
of failure pf as a function of  for a range of input
point coefficients of variation, with the point mean fixed at
c = 0.25. The probability of failure is defined, as before,

0.1 0
by p[c < 0.17], but this time the calculation is based on the
finite-element locally averaged properties, cA and cA from 2 3 4 5 6 7 89 2 3 4 5 6 7 89
Eqs. 13.13. The Figure clearly shows two tails to the results, 101 100 101
uc
with pf 1 as  0 for all vc > 1.0783, and pf 0 as
 0 for all vc < 1.0783. The horizontal line at pf = 0.5 Figure 13.14 Probability of failure versus coefficient of varia-
is given by vc = 1.0783, which is the special value of tion based on finite-element locally geometrically averaged prop-
the coefficient of variation that causes the median of c erties; the mean is fixed at c = 0.25.
to have value c = 0.17. Recalling Table 13.1, this is the
critical value of c that would give FS = 1 in the test slope.
Higher values of vc lead to c < 0.17 and a tendency for interest to note the step function corresponding to  = 0
pf 1 as  0. Conversely, lower values of vc lead to when pf changes suddenly from zero to unity.
c > 0.17 and a tendency for pf 0. Figure 13.14 shows It should be emphasized that the results presented in this
the same data plotted the other way round with vc along section involved no finite-element analysis and were based
the abscissa. This Figure clearly shows the full influence of solely on an SRV approach with statistical properties based
spatial correlation in the range 0  < . All the curves on finite-element locally geometrically averaged properties
cross over at the critical value of vc = 1.0783, and it is of based on a typical finite element of the mesh in Figure 13.8.

13.2.9 Results of RFEM Analyses


0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

In this section, the results of full nonlinear RFEM analyses


with Monte Carlo simulations are described, based on a
uc = 8 range of parametric variations of c , vc , and .
uc = 4 In the elasto-plastic RFEM approach, the failure mech-
uc = 2
uc = 1.0783 anism is free to seek out the weakest path through the
pf

uc = 1 soil. Figure 13.15 shows two typical random field realiza-


uc = 0.5
uc = 0.25 tions and the associated failure mechanisms for slopes with
 = 0.5 and  = 2. The convoluted nature of the failure
mechanisms, especially when  = 0.5, would defy anal-
ysis by conventional slope stability analysis tools. While
the mechanism is attracted to the weaker zones within the
2 4 68 2 4 68 2 4 68 2 4 68 2 4 68
slope, it will inevitably pass through elements assigned
102 101 100 101 102 103
many different strength values. This weakest path determi-

nation, and the strength averaging that goes with it, occurs
Figure 13.13 Probability of failure versus spatial correlation quite naturally in the finite-element slope stability method
length based on finite-element locally geometrically averaged and represents a very significant improvement over tradi-
properties; the mean is fixed at c = 0.25. tional limit equilibrium approaches to probabilistic slope
390 13 SLOPE STABILITY

= 0.5 clearly indicates two branches, with the probability of fail-


ure tending to unity or zero for higher and lower values
of vc , respectively. This behavior is qualitatively simi-
lar to that observed in Figure 13.13, in which an SRV
approach was used to predict the probability of failure
based solely on finite-element locally averaged properties.
Figure 13.17 shows the same results as Figure 13.16, but
= 2.0 plotted the other way round with the coefficient of vari-
ation along the abscissa. Figure 13.17 also demonstrates
that when  becomes large, corresponding approximately
to an SRV approach with no local averaging, the proba-
bility of failure is overestimated (conservative) when the
coefficient of variation is relatively small and underesti-
Figure 13.15 Typical random-field realizations and deformed mated (unconservative) when the coefficient of variation
mesh at slope failure for two different spatial correlation lengths. is relatively high. Figure 13.17 also demonstrates that the
Light zones are weaker. SRV approach described earlier in the section, which gave
pf = 0.28 corresponding to c = 0.25 and vc = 0.5 with
no local averaging, is indeed pessimistic. The RFEM results
show that the inclusion of spatial correlation and local aver-
stability, in which local averaging, if included at all, has
aging in this case will always lead to a smaller probability
to be computed over a failure mechanism that is preset
of failure.
by the particular analysis method (e.g., a circular failure
Comparison of Figures 13.13 and 13.14 with Figures
mechanism when using Bishops method).
13.16 and 13.17 highlights the influence of the finite-
Fixing the point mean strength at c = 0.25, Fig-
element approach to slope stability, where the failure
ures 13.16 and 13.17 show the effect of the spatial cor-
mechanism is free to locate itself optimally within the mesh.
relation length  and the coefficient of variation vc on the From Figures 13.14 and 13.17, it is clear that the weak-
probability of failure for the test problem. Figure 13.16 est path concept made possible by the RFEM approach
1
1

0.9
0.9

0.8
0.8

0.7
0.7

0.6
0.6

uc = 0.25
0.5
0.5

pf
pf

uc = 0.5
uc = 1 pf = 0.38 = 0.5
0.4
0.4

uc = 2 =1
uc = 4 =2
=4
0.3
0.3

uc = 8
=8
c = 0.65
0.2
0.2

0.1
0.1

0
0

2 3 4 5 6 7 89 2 3 4 5 6 7 89 2 3 4 5 6 7 89 2 3 4 5 6 7 89
101 100 101 101 100 101
c

Figure 13.16 Probability of failure versus spatial correlation Figure 13.17 Probability of failure versus coefficient of varia-
length from RFEM; the mean is fixed at c = 0.25. tion from RFEM; the mean is fixed at c = 0.25.
PROBABILISTIC SLOPE STABILITY ANALYSIS 391
has resulted in the crossover point falling to lower values In all cases, as  increases, the RFEM and the locally
of both vc and pf . With only finite-element local averag- averaged solutions converge on the SRV solution corre-
ing, the crossover occurred at vc = 1.0783, whereas by the sponding to  = with no local averaging. The pf = 0.28
RFEM it occurred at vc 0.65. In terms of the probabil- value, corresponding to vc = 0.5, and discussed earlier in
ity of failure with only finite-element local averaging, the the section, is also indicated in Figure 13.18.
crossover occurred at pf = 0.5, whereas by the RFEM it All of the above results and discussion in this section so
occurred at pf 0.38. The RFEM solutions show that the far were applied to the test slope from Figure 13.1 with the
SRV approach becomes unconservative over a wider range mean strength fixed at c = 0.25 corresponding to a factor
of vc values than would be indicated by finite- element local of safety (based on the mean) of 1.47. In the next set of
averaging alone. results c is varied while vc is held constant at 0.5. Figure
Figure 13.18 gives a direct comparison between Figures 13.19 shows the relationship between FS (based on the
13.13 and 13.16, indicating clearly that for higher values mean) and pf assuming finite-element local averaging only,
of vc , RFEM always gives a higher probability of failure and Figure 13.20 shows the same relationship as computed
than when using finite- element local averaging alone. This using RFEM.
is caused by the weaker elements in the distribution domi- Figure 13.19, based on finite-element local averaging
nating the strength of the slope and the failure mechanism only, shows the full range of behavior for 0  < .
seeking out the weakest path through the soil. The figure shows that  only starts to have a significant
At lower values of vc , the locally averaged results tend to influence on the FS vs. pf relationship when the correlation
length becomes significantly smaller than the slope height
overestimate the probability of failure and give conservative
( << 1). The step function in which pf jumps from zero
results compared with RFEM. In this case the stronger
to unity occurs when  = 0 and corresponds to a local
elements of the slope are dominating the solution, and the
average having zero variance. In this limiting case, the
higher median combined with the bunching up of the
local average of the soil is deterministic, yielding a constant
locally averaged solution at low values of  means that
strength everywhere in the slope. With vc = 0.5, the critical
potential failure mechanisms cannot readily find a weak
value of mean shear strength that would give cA = c =
path through the soil.
0.17 is easily shown by Eq. 13.14 to be c = 0.19, which
corresponds to an FS = 1.12. For higher values of ,
1.1
1
0.9

uc = 8
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
0.8

FS = 1.12

uc = 4
0.7

uc = 2 =0
0.6

= 0.0625
= 0.125
0.5

uc = 1 = 0.25
pf

= 0.5
pf
0.4

=1
=
0.3

uc = 0.5
(pf 0.28)
0.2
0.1

uc = 0.25
0

2 3 4 5 6 7 89 2 3 4 5 6 7 89
101
0.1

100 101
0.8 1 1.2 1.4 1.6 1.8 2
FS
Figure 13.18 Comparison of the probabilities of failure pre-
dicted by RFEM and by finite-element local geometric averaging Figure 13.19 Probability of failure versus FS (based on mean)
alone; the curves which include points come from the random using finite-element local geometric averaging only for test slope;
finite-element method; the mean is fixed at c = 0.25. the coefficient of variation is fixed at vc = 0.5.
392 13 SLOPE STABILITY

the relationship between FS and pf is quite bunched up

1.1
and generally insensitive to . For example, there is little RFEM

1
difference between the curves corresponding to  = and
 = 0.5. It should also be observed from Figure 13.19 that

0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9


for FS > 1.12, failure to account for local averaging by
Finite-element
assuming  = is conservative, in that the predicted pf is local averaging
higher than it should be. When FS < 1.12, however, failure
to account for local averaging is unconservative.
Figure 13.20 gives the same relationships as computed

pf
using RFEM. By comparison with Figure 13.19, the RFEM
results are more spread out, implying that the probability
of failure is more sensitive to the spatial correlation length
. Of greater significance is that the crossover point has
again shifted by RFEM as it seeks out the weakest path
through the slope. In Figure 13.20, the crossover occurs
at FS 1.37, which is significantly higher and of greater

0
practical significance than the crossover point of FS 1.12
by finite-element local geometric averaging alone. The 0.1
0.8 1 1.2 1.4 1.6 1.8 2
theoretical line corresponding to  = is also shown in FS
this plot. From a practical viewpoint, the RFEM analysis
indicates that failure to properly account for local averaging Figure 13.21 Comparison of probabilities of failure versus FS
is unconservative over a wider range of factors of safety (based on mean) using finite-element local geometric averaging
than would be the case by finite-element local averaging alone with RFEM for test slope; vc = 0.5 and ln c = 0.5.
alone. To further highlight this difference, the particular
results from Figures 13.19 and 13.20 corresponding to  =
0.5 (spatial correlation length equal to half the embankment 13.2.10 Summary
height) have been replotted in Figure 13.21. The section has investigated the probability of failure of
a cohesive slope using both simple and more advanced
probabilistic analysis tools. The simple approach treated
1.1

the strength of the entire slope as a single random variable,


ignoring spatial correlation and local averaging. In the sim-
1

ple studies, the probability of failure was estimated as the


0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

probability that the shear strength would fall below a criti-


cal value based on a lognormal probability density function.
= 0.5 These results led to a discussion on the appropriate choice
=1 of a design shear strength value suitable for determinis-
=2
=4 tic analysis. Two factorization methods were proposed that
=8 were able to bring the probability of failure and the FS
pf

=
more into line with practical experience.
pf = 0.35
The second half of the section implemented the RFEM on
the same test problem. The nonlinear elasto-plastic analyses
with Monte Carlo simulation were able to take full ac-
count of spatial correlation and local averaging and observe
FS = 1.37

their impact on the probability of failure using a parametric


approach. The elasto-plastic finite-element slope stability
0

method makes no a priori assumptions about the shape or


0.1

0.8 1 1.2 1.4 1.6 1.8 2


location of the critical failure mechanism and, therefore,
FS offers very significant benefits over traditional limit equi-
librium methods in the analysis of spatially variable soils.
Figure 13.20 Probability of failure versus FS (based on mean) In the elasto-plastic RFEM, the failure mechanism is free to
using RFEM for test slope; the coefficient of variation is fixed at seek out the weakest path through the soil, and it has been
vc = 0.5. shown that this generality can lead to higher probabilities
SLOPE STABILITY RELIABILITY MODEL 393
of failure than could be explained by finite-element local similar, except that the horizontal length of the slope is H
averaging alone. rather than 2H .
In summary, simplified probabilistic analysis in which The soil is represented by a random spatially varying
spatial variability is ignored by assuming perfect correla- undrained cohesion field cu (x) which is assumed to be
tion can lead to unconservative estimates of the probability lognormally distributed, where x is the spatial position.
of failure. This effect is most pronounced at relatively The cohesion has mean cu and standard deviation cu and
low factors of safety (Figure 13.20) or when the coef- is assumed to have an exponentially decaying (Markovian)
ficient of variation of the soil strength is relatively high correlation structure:
(Figure 13.18).
ln cu ( ) = e 2| |/ln cu (13.17)

13.3 SLOPE STABILITY RELIABILITY MODEL where is the distance between two points in the field. Note
that the correlation structure has been assumed isotropic
The failure prediction of a soil slope has been a long- in this study. The use of an anisotropic correlation is
standing geotechnical problem and one which has attracted straightforward, within the framework developed here, but
a wide variety of solutions. Traditional approaches to the is considered a site-specific extension. In this section it
problem generally involve assuming that the soil slope is is desired to investigate the stochastic behavior of slope
homogeneous (spatially constant) or possibly layered, and stability for the simpler isotropic case, leaving the effect of
techniques such as Taylors (1937) stability coefficients for anisotropy for the reader.
frictionless soils, the method of slices, and other more
The correlation function has a single parameter, ln cu ,
general methods involving arbitrary failure surfaces have
the correlation length. Because cu is assumed to be log-
been developed over the years. The main drawback to
normally distributed, its logarithm, ln cu , is normally dis-
these methods is that they are not able to easily find the
tributed. In this study, the correlation function is measured
critical failure surface in the event that the soil properties
relative to the underlying normally distributed field. Thus,
are spatially varying.
ln cu ( ) gives the correlation coefficient between ln cu (x)
In the realistic case where the soil properties vary ran-
and ln cu (x ) at two points in the field separated by the dis-
domly in space, the slope stability problem is best captured
tance = |x x |. In practice, the parameter ln cu can be
via a nonlinear finite-element model which has the dis-
estimated from spatially distributed cu samples by using
tinct advantage of allowing the failure surface to seek out
the logarithm of the samples rather than the raw data them-
the path of least resistance, as pointed out in the previous
selves. If the actual correlation between points in the cu
section. In this section such a model is employed, which,
field is desired, the following transformation can be used
when combined with a random-field simulator, allows the
(Vanmarcke, 1984):
realistic probabilistic evaluation of slope stability (Fenton
and Griffiths, 2005c). This work builds on the previous exp{ln cu ( )ln2 cu } 1
section, which looked in some detail at the probability of cu ( ) = (13.18)
exp{ln2 cu } 1
failure of a single slope geometry. Two slope geometries
are considered in this section, one shallower with a 2 : 1 The spatial correlation length can be nondimensionalized by
gradient and the other steeper with a 1 : 1 gradient. Both dividing it by the slope height H as was done in Eq. 13.4:
slopes are assumed to be composed of undrained clay, with ln cu
u = 0, of height H with the slope resting on a foundation = (13.19)
layer, also of depth H . The finite-element mesh for the 2 : 1 H
gradient slope is shown in Figure 13.22. The 1 : 1 slope is Thus, the results given here can be applied to any size
problem, so long as it has the same slope and same overall
bedrock depthslope height ratio D. The standard deviation
2H 2H 2H cu may also be expressed in terms of the dimensionless
coefficient of variation
Unit weight g H c
vc = u (13.20)
cu
H If the mean and variance of the underlying ln cu field are
desired, they can be obtained through the transformations
 
Figure 13.22 Mesh used for stability analysis of 2 : 1 gradient ln2 cu = ln 1 + vc2 , ln cu = ln(cu ) 12 ln2 cu
slope. (13.21)
394 13 SLOPE STABILITY

By using Monte Carlo simulation, where the soil slope is continuum as well as the way that soil samples are typi-
simulated and analyzed by the finite-element method re- cally taken and tested in practice, that is, as local averages.
peatedly, estimates of the probability of failure are obtained Regarding the discretization of random fields for use in
over a range of soil statistics. The failure probabilities finite-element analysis, Matthies et al. (1997, p. 294) makes
are compared to those obtained using a harmonic aver- the following comment: One way of making sure that the
age of the cohesion field employed in Taylors stability stochastic field has the required structure is to assume that
coefficient method, and very good agreement is found. The it is a local averaging process, referring to the conversion
study indicates that the stability of a spatially varying soil of a nondifferentiable to a differentiable (smooth) stochastic
slope is well modeled using a harmonic average of the soil process. Matthies further goes on to say that the advantage
properties. of the local average representation of a random field is that
it yields accurate results even for rather coarse meshes.
13.3.1 Random Finite-Element Method Figure 13.23 illustrates two possible realizations arising
from the RFEM for the 2 : 1 slopesimilar results were
The slope stability analyses use an elastic-perfectly plas- observed for the 1 : 1 slope. In this figure, dark regions
tic stressstrain law with a Tresca failure criterion. Plastic correspond to stronger soil. Notice how convoluted the
stress redistribution is accomplished using a viscoplastic failure region is, particularly at the smaller correlation
algorithm which uses 8-node quadrilateral elements and length. It can be seen that the slope failure involves the
reduced integration in both the stiffness and stress redis- plastic deformation of a region around a failure surface
tribution parts of the algorithm. The theoretical basis of the which undulates along the weakest path. Clearly, failure
method is described more fully in Chapter 6 of the text is more complex than just a rigid circular region sliding
by Smith and Griffiths (2004). The method is discussed in along a clearly defined interface, as is typically assumed.
more detail in the previous section.
In brief, the analyses involve the application of grav-
13.3.2 Parametric Studies
ity loading and the monitoring of stresses at all the Gauss
points. If the Tresca criterion is violated, the program at- To keep the study nondimensional, the soil strength is
tempts to redistribute those stresses to neighboring elements expressed in the form of a dimensionless shear strength:
that still have reserves of strength. This is an iterative pro- cu
cess which continues until the Tresca criterion and global c= (13.22)
H
equilibrium are satisfied at all points within the mesh under
which, if cu is random, has mean
quite strict tolerances.
c
In this study,failure is said to have occurred if, for any c = u (13.23)
given realization, the algorithm is unable to converge within H
500 iterations (see Figure 13.10). Following a set of 2000 where is the unit weight of the soil, assumed in this
realizations of the Monte Carlo process the probability of study to be deterministic. In the 2 : 1 slope case where the
failure is simply defined as the proportion of these realiza- cohesion field is assumed to be everywhere the same and
tions that required 500 or more iterations to converge.
The RFEM combines the deterministic finite -element
analysis with a random-field simulator, which, in this study, = 0.5
is the LAS discussed in Section 6.4.6. The LAS algorithm
produces a field of random element values, each represent-
ing a local average of the random field over the element
domain, which are then mapped directly to the finite el-
ements. The random elements are local averages of the
log-cohesion, ln cu , field. The resulting realizations of the
log-cohesion field have correlation structure and variance
correctly accounting for local averaging over each element. = 2.0
Much discussion of the relative merits of various methods
of representing random fields in finite-element analysis has
been carried out in recent years (see, e.g., Li and Der Ki-
ureghian, 1993). While the spatial averaging discretization
of the random field used in this study is just one approach
to the problem, it is appealing in the sense that it reflects Figure 13.23 Two typical failed random-field realizations. Low-
the simplest idea of the finite-element representation of a strength regions are light.
SLOPE STABILITY RELIABILITY MODEL 395
equal to cu , a value of c = 0.173 corresponds to a factor [namely, cu = cu (x)] so that c becomes random. The first
of safety FS = 1.0, which is to say that the slope is on the issue can be solved by finding some representative or
verge of failure. For the 1 : 1 slope, c = 0.184 corresponds equivalent value of cu , which will be referred to here as
to a factor of safety FS = 1.0. Both of these values were cu , such that the stability coefficient method still holds for
determined by finding the deterministic value of cu needed the slope. That is, cu would be the cohesion of a uniform
to just achieve failure in the finite-element model, bearing soil such that it has the same factor of safety as the real
in mind that the failure surface cannot descend below the spatially varying soil.
base of the model. These values are almost identical to what The question now is: How should this equivalent soil
would be identified using Taylors charts (Taylor, 1937), cohesion value be defined? First of all, each soil realization
although as will be seen later, small variations in the choice will have a different value of cu , so that Eq. 13.24 is still
of the critical values of c can result in significant changes a function of a random quantity, namely
in the estimated probability of slope failure, particularly for cu
larger factors of safety. c= (13.25)
H
This study considers the following values of the input
statistics. For the 2 : 1 slope, c is varied over the following If the distribution of cu is found, the distribution of c can be
derived. The failure probability of the slope then becomes
values:
equal to the probability that c is less than the Taylor critical
c = 0.15, 0.173, 0.20, 0.25, 0.30
value ccrit .
and over This line of reasoning suggests that cu should be de-
c = 0.15, 0.184, 0.20, 0.25, 0.30 fined as some sort of average of cu over the soil domain
where failure is occurring. Three common types of averages
for the 1 : 1 slope. For the normalized correlation length  present themselves, as discussed in Section 4.4:
and coefficient of variation vc , the following ranges were
investigated: 1. Arithmetic Average: The arithmetic average over
some domain, A, is defined as
 = 0.10, 0.20, 0.50, 1.00, 2.00, 5.00, 10.0

1
n
1
Xa = cui = cu (x) d x (13.26)
vc = 0.10, 0.20, 0.50, 1.00, 2.00, 5.00 n A A
i =1
For each set of the above parameters, 2000 realizations for the discrete and continuous cases, where the do-
of the soil field were simulated and analyzed, from which main A is assumed to be divided up into n samples in
the probability of slope failure was estimated. This section the discrete case. The arithmetic average weights all
concentrates on the development of a failure probability of the values of cu equally. In that the failure surface
model, using a harmonic average of the soil, and compares seeks a path through the weakest parts of the soil, this
the simulated probability estimates to those predicted by form of averaging is not deemed to be appropriate for
the harmonic average model. this problem.
2. Geometric Average: The geometric average over
13.3.3 Failure Probability Model some domain, A, is defined as
 n 1/n 

In Taylors stability coefficient approach to slope stability  1
(Taylor, 1937), the coefficient Xg = cui = exp ln cu (x) d x
A A
cu i =1
c= (13.24) (13.27)
H The geometric average is dominated by low values
assumes that the soil is completely uniform, having co- of cu and, for a spatially varying cohesion field,
hesion equal to cu everywhere. This coefficient may then will always be less than the arithmetic average. This
be compared to the critical coefficient obtained from Tay- average potentially reflects the reduced strength as
lors charts to determine if slope failure will occur or not. seen along the failure path and has been found by
For the slope geometry studied here, slope failure will oc- the authors (Fenton and Griffiths, 2002, 2003) to
cur if c < ccrit where ccrit = 0.173 for the 2 : 1 slope and well represent the bearing capacity and settlement
ccrit = 0.184 for the 1 : 1 slope. of footings founded on spatially random soils. The
In the case where cu is randomly varying in space, geometric average is also a natural average of the
two issues present themselves. First of all Taylors method lognormal distribution since an arithmetic average of
cannot be used on a nonuniform soil, and, second, Eq. 13.24 the underlying normally distributed random variable,
now includes a random quantity on the right-hand side ln cu , leads to the geometric average when converted
396 13 SLOPE STABILITY

back to the lognormal distribution. Thus, if cu is comparing Figure 13.23 to 13.24, that the assumed averag-
lognormally distributed, its geometric local average ing domain of Figure 13.24 is smaller than the deformed
will also be lognormally distributed with the median regions seen in Figure 13.23. A general prescription for the
preserved. size of the averaging domain is not yet known, although
3. Harmonic Average: The harmonic average over some it should capture the approximate area of the soil involved
domain, A, is defined as in resisting the slope deformation. The area assumed in
 n 1 
 Figure 13.24 is to be viewed as an initial approximation
1 1 1 d x 1 which, as will be seen, yields surprisingly good results. It
Xh = = (13.28)
n cui A A cu (x) is recognized that the true average will be of the minimum
i =1
soil strengths within a roughly circular bandpresumably
This average is even more strongly influenced by
the area of this band is on average approximated by the
small values than is the geometric average. In general,
area shown in Figure 13.24.
for a spatially varying random field, the harmonic
With an assumed averaging domain, A = w h, the
average will be smaller than the geometric average,
geometric average leads to the following definition for cu :
which in turn is smaller than the arithmetic average. 

Unfortunately, the mean and variance of the harmonic 1
cu = Xg = exp ln cu (x) d x (13.30)
average, for a spatially correlated random field, are not A A
easily found. which, if cu is lognormally distributed, is also lognormally
distributed. The resulting coefficient
Putting aside for the moment the issue of how to compute
cu
the equivalent undrained cohesion, cu , the size of the c= (13.31)
averaging domain must also be determined. This should H
approximately equal the area of the soil which fails during is then also lognormally distributed with mean and variance
a slope subsidence. Since the value of cu changes only
slowly with changes in the averaging domain, only an ln c = ln cu ln( H ) (13.32a)
approximate area need be determined. The area selected ln2 c = ln2 cu = (w, h)ln2 cu (13.32b)
in this study is a parallelogram, as shown in Figure 13.24,
having slope length equal to the length of the slope and The function (w, h) is the so-called variance function,
horizontal surface length equal to H . For the purposes of which lies between 0 and 1, and gives the amount that
computing the average, it is further assumed that this area the variance of a local average is reduced from the point
can be approximated by a rectangle of dimension w h value. It is formally defined as the average of correlations
(averages over rectangles are generally easier to compute). between every pair of points in the averaging domain:

Thus, a rectangular w h area is used to represent a 1


roughly circular band (on average) within which the soil (w, h) = 2 ( ) d d (13.33)
A A A
is failing in shear. Solutions to this integral, albeit sometimes approximate,
In this study, the values of w and h are taken to be exist for most common correlation functions. Alternatively,
H the integral can be calculated accurately using a numerical
w= , h = H sin (13.29)
sin method such as Gauss quadrature. See Appendix C for more
such that w h = H 2 , where is the slope angle (26.6 for details.
the 2 : 1 slope and 45 for the 1 : 1 slope). It appears, when The probability of failure pf can now be computed by
assuming that Taylors stability coefficient method holds
when using this equivalent value of cohesion, namely by
H computing
 
ln ccrit ln c
w pf = P [c < ccrit ] =  (13.34)
A H ln c
h where the critical stability coefficient for the 2 : 1 slope is
ccrit = 0.173 and for the 1 : 1 slope is ccrit = 0.184;  is
H the cumulative distribution function for the standard nor-
mal. Unfortunately, the geometric average for cu leads to
predicted failure probabilities which significantly under-
2H 2H 2H
estimate the probabilities determined via simulation, and
Figure 13.24 Assumed averaging domain (1 : 1 slope is similar). changes in the averaging domain size does not particularly
SLOPE STABILITY RELIABILITY MODEL 397
improve the prediction. This means that the soil strength as The procedure to estimate the mean and variance of
seen by the finite-element model is even lower, in gen- the harmonic average cu for each parameter set (c , vc ,
eral, than that predicted by the geometric average. Thus, the and ) considered in this study involves (a) generating
geometric average was abandoned as the correct measure a large number of random cohesion fields, each of di-
for cu . mension w h, (b) computing the harmonic average of
Since the harmonic average yields values which are even each using Eq. 13.28, and (c) estimating the mean and
lower than the geometric average, the harmonic average variance of the resulting set of harmonic averages. Us-
over the same domain, A = w h, is now investigated as ing 5000 random-field realizations, the resulting estimates
representative of cu , namely for the mean and standard deviation of ln Xh are shown

 in Figure 13.26 for random fields with mean 1.0. Since
1 d x 1 cu is assumed to be (at least approximately) lognormally
cu = Xh = (13.35)
A A cu (x) distributed, having parameters ln cu and ln cu , the mean
Unfortunately, so far as the authors are aware, no relatively and standard deviation of the logarithm of the harmonic
simple expressions exist for the moments of cu , as defined averages are shown in Figure 13.26 for the two slopes con-
above, for a spatially correlated random field. The authors sidered. Of note in Figure 13.26 is the fact that there is
are continuing research on this problem but, for the time virtually no difference in the mean and standard deviation
being, these moments can be obtained by simulation. It for the 2 : 1 and 1 : 1 slopes, even though the averaging
may seem questionable to be developing a probabilistic regions have quite different shapes. Admittedly the two
model with the nominal goal of eliminating the necessity averaging regions have the same area, but this only slow
of simulation, when that model still requires simulation. change in harmonic average statistics with averaging di-
However, the moments of the harmonic mean can be arrived mension has been found also to be true of changing areas.
at in a small fraction of the time taken to perform the This implies that the accurate determination of the averag-
nonlinear slope stability simulation. ing area is not essential to the accuracy of failure probability
In order to compute probabilities using the statistics predictions.
of cu , it is necessary to know the distribution of c = Given the results of Figure 13.26, the slope failure
cu /( H ). For lognormally distributed cu , the distribution probability can now be computed as in Eq. 13.34:
 
of the harmonic average is not simple. However, since ln ccrit ln c
cu is strictly nonnegative (cu 0), it seems reasonable pf = P [c < ccrit ] =  (13.36)
ln c
to suggest that cu is at least approximately lognormal. A
histogram of the harmonic averages obtained in the case except that now the mean and standard deviation of ln c are
where vc = 0.5 and  = 0.5 is shown in Figure 13.25, computed using the harmonic mean results of Figure 13.26
along with a fitted lognormal distribution. The p-value suitably scaled for the actual value of cu / H as follows:
for the chi-Square goodness-of-fit test is 0.44, indicating  
cu
that the lognormal distribution is very reasonable, as also ln c = ln + ln X h = ln(c ) + ln X h (13.37a)
indicated by the plot. Similar results were obtained for other H
parameter values. ln c = ln X h (13.37b)
where ln X h and ln X h are read from Figure 13.26, given
the correlation length and coefficient of variation.
Frequency density Figure 13.27 shows the predicted failure probabilities
3

mln X = 0.204, sln X = 0.187, p-value = 0.44


versus the failure probabilities obtained via simulation over
all parameter sets considered. The agreement is remarkably
good, considering the fact that the averaging domain was
2
fXh(x)

rather arbitrarily selected, and there was no a priori evi-


dence that the slope stability problem should be governed
1

by a harmonic average. The results of Figure 13.27 indicate


that the harmonic average gives a good probabilistic model
of slope stability.
0

0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 There are a few outliers in Figure 13.27 where the pre-
x
dicted failure probability considerably overestimates that
Figure 13.25 Histogram of harmonic averages along with fitted obtained via simulation. For the 2 : 1 slope, these out-
lognormal distribution. liers correspond to the cases where (1) c = 0.3, vc = 1.0,
398 13 SLOPE STABILITY

0
1

1
mln Xh

mln Xh
q = 0.1 q = 0.10
2

2
q = 0.2 q = 0.20
q = 0.5 q = 0.50
q = 1.0 q = 1.00
q = 2.0 q = 2.00
3

3
q = 5.0 q = 5.00
q = 10.0 q = 10.00
4

4
2 4 6 8 2 4 6 8 2 4 6 8 2 4 6 8
101 100 101 101 100 101
s/m s/m
2

q = 0.10 q = 0.10
q = 0.20 q = 0.20
q = 0.50 q = 0.50
q = 1.00 q = 1.00
1.5

1.5

q = 2.00 q = 2.00
q = 5.00 q = 5.00
q = 10.00 q = 10.00
sln Xh

sln Xh
1

1
0.5

0.5
0

2 4 6 8 2 4 6 8 2 4 6 8 2 4 6 8
101 100 101 101 100 101
s/m s/m
(a) (b)

Figure 13.26 Mean and standard deviation of log-harmonic averages estimated from 5000
simulations: (a) 2 : 1 cohesive slope; (b) 1 : 1 cohesive slope.

and  = 0.1 (simulated probability is 0.047 versus pre- For example, the worst case seen in Figure 13.27a has
dicted probability of 0.86) and (2) c = 0.3, vc = 1.0, predicted values of
and  = 0.2 (simulated probability is 0.31 versus pre-
dicted probability of 0.74). Both cases correspond to the ln c = ln(c ) + ln X h = ln(0.3) 0.66 = 1.864
largest FS considered in the study (c = 0.3 gives an
ln c = ln X h = 0.10
FS = 1.77 in the uniform soil case). Also the small cor-
relation lengths yield the smallest values of ln c which, in The predicted failure probability is thus
turn, implies that the cumulative distribution function of  
ln c increases very rapidly over a small range. Thus, slight ln 0.173 + 1.864
P [c < 0.173] =  = (1.10)
errors in the estimate of ln c makes for large errors in the 0.10
probability. = 0.86
SLOPE STABILITY RELIABILITY MODEL 399
1

1
= 0.1 = 0.1
= 0.2 = 0.2
= 0.5 = 0.5
0.8

0.8
=1 =1
=2 =2
=5 =5
Simulated P[failure]

Simulated P[failure]
= 10 = 10
0.6

0.6
0.4

0.4
0.2

0.2
0

0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Predicted P[failure] Predicted P[failure]
(a) (b)

Figure 13.27 Simulated failure probabilities versus failure probabilities predicted using a har-
monic average of cu over domain w h: (a) 2 : 1 cohesive slope; (b) 1 : 1 cohesive slope.

As mentioned, a relatively small error in the estimation of For all other results, especially where the FS is closer
ln c can lead to a large change in probability. For example, to 1.0 (c < 0.3), the harmonic average model leads to
if ln c was 1.60 instead of 1.864, a 14% change, then very good estimates of failure probability with somewhat
the predicted failure probability changes significantly to more scatter seen for the 1 : 1 slope. The increased scatter
  for the 1 : 1 slope is perhaps as expected since the steeper
ln 0.173 + 1.6 slope leads to a larger variety of critical failure surfaces. In
P [c < 0.173] =  = (1.54)
0.10 general, for both slopes the predicted failure probability
= 0.062 is seen to be conservative at small failure probabilities,
slightly overestimating the failure probability.
which is about what was obtained via simulation. The
conclusion drawn from this example is that small errors
in the estimation of ln c or, equivalently, in ccrit can lead 13.3.4 Summary
to large errors in the predicted slope failure probability if
the standard deviation of ln c is small. The latter occurs This study investigates the failure probabilities of two
for small correlation lengths, . In most cases for small undrained clay slopes, one with gradient 2 : 1 and the
values of  the failure probability tends to be either close other with gradient 1 : 1. The basic idea of the section
to zero (vc < 1.0) or close to 1.0 (vc > 1.0), in which case is that the Taylor stability coefficients are still useful if
the predicted and simulated probabilities are in much better an equivalent soil property can be found to represent
agreement. That is, the model shows very good agreement the spatially random soil. It was found that a harmonic
with simulation for all but the case where a large FS is average of the soil cohesion over a region of dimension
combined with a small correlation length and intermediate H 2 (sin 1/ sin ) = H 2 yields an equivalent stabil-
coefficient of variation (vc  1.0). This means that the ity number with an approximately lognormal distribution
selected harmonic average model is not the best predictor that quite well predicts the probability of slope failure.
in the region where the cumulative distribution is rapidly The harmonic average was selected because it is domi-
increasing. However, in these cases, the predicted failure nated by low-strength regions appearing in the soil slope,
probability is overestimated, which is at least conservative. which agrees with how the failure surface will seek out the
400 13 SLOPE STABILITY

low-strength areas. The dimension of the averaging region how soil samples are treated. In particular, the study sug-
was rather arbitrarily selectedthe equivalent stability co- gests that the reliability of an existing slope is best estimated
efficient mean and variance is only slowly affected by by sampling the soil at a number of locations and then us-
changes in the averaging region dimensionbut is believed ing a harmonic average of the sample values to estimate
to reasonably approximate the area of the average slope the soils equivalent cohesion. Most modern geotechnical
failure band. codes suggest that soil design properties be taken as cau-
An important practical conclusion arising from the fact tious estimates of the meanthe harmonic average, being
that soil slopes appear to be well characterized by a har- governed by low-strength regions, is considered by the au-
monic average of soil sample values, rather than by an thors to be such a cautious estimate for slope stability
arithmetic average, as is traditionally done, has to do with calculations.

You might also like