Sunteți pe pagina 1din 29

Chapter 8

Probabilistic assessment of randomly


heterogeneous soil slopes

8.1 SPATIAL VARIABILITY OF SOILS

Soils are geological materials formed by weathering processes, transported by physical


means to their present locations. They have been subject to various stresses, pore fluids,
and physical and chemical changes. Thus, it is not surprising that the physical properties
vary from place to place, even in one soil stratum. Researchers have identified and classi-
fied major components of uncertainties associated with the estimation of soil properties
(Vanmarcke, 1977; Tang, 1984; Phoon and Kulhawy, 1999; Baecher and Christian, 2003;
Zhang etal., 2003). Vanmarcke (1977) presented basic concepts and methods for describ-
ing the spatial variability of a soil stratum within random field framework. Three major
sources of uncertainties were identified for stochastic modeling of soil profiles. The first is
natural heterogeneity or in situ variability of the soil, which is caused by variation in mineral
composition, varying depths of strata during soil formation and stress history. The second
is attributed to limited availability of information about subsurface conditions. The third
is measurement errors due to sample disturbance, test imperfections, human factors, and
estimating soil properties through correlation with index properties. Kulhawy (1992) classi-
fied three primary sources of geotechnical uncertainties: inherent variability, measurement
uncertainties, and transformation uncertainties (Figure8.1), which are basically the same
as those defined by Vanmarcke (1977). The uncertainties caused by equipment, operator,
random test effects, and statistical uncertainties were all included in measurement errors.
The transformation uncertainties meant the uncertainty introduced when field or laboratory
measurements are transformed into design soil properties using empirical or other correla-
tion models.
The spatial variability in a soil profile may be either continuous or discrete (JCSS 2002).
The continuous spatial variability may be characterized by an average trend and continu-
ous fluctuations around the average trend (Figure8.2). Usually, this type of variability is
modeled as a stationary random field. Soils may be mixed with dislocations such as faults
and lenses depending on the geological history. This type of spatial variability is defined
as discrete spatial variability. Though local of nature, these features may have a significant
influence on the performance of a geotechnical engineering structure. In this book, we will
focus on the continuous spatial variability.
The effect of inherent spatial variability of soil properties on the performance of geotech-
nical works has received considerable attention in recent years. Some researchers have con-
sidered the spatial variability for the problem of flow in unsaturated soil and slope stability.
Zhang (1999), Tartakovsky et al. (2003), Yang et al. (2004), Lu and Zhang (2007), Li et al.
(2009), Mousavi Nezhad et al. (2011), Cho (2012), Le et al. (2012), and Meftah et al. (2012)
among others, analyzed unsaturated flow in heterogeneous soils with spatially distributed uncer-
tain hydraulic parameters. Cho (2007), Jiang et al. (2014a,b) Le (2014), Li et al. (2014; 2015),

299
300 Rainfall-Induced Soil Slope Failure

In situ Transformation Estimated soil


Soil
measurement model property

Inherent Data Statistical Model


soil variability scatter uncertainty uncertainty

Inherent Measurement
soil variability error

Uncertainties in estimation of soil properties. (From Phoon, K. K., and Kulhawy, F. H., Geotechnical
Figure8.1
Journal, 36, 612624, 1999. With permission.)
Downloaded by [Monash University Library] at 23:45 02 August 2017

Ground surface Ground surface

L1 Soil layer 1 L1 Soil layer 1

Trend

L2 Soil layer 2
L2 Soil layer 2

L3 Soil layer 3
L3 Soil layer 3

(a) (b)

Figure8.2Inherent spatial variability of a soil property (a) without a trend and (b) with a trend.

and Li and Jiang (2015) assessed the effect of inherent variability on slope stability. Srivastava
etal. (2010), Santoso et al. (2011), and Zhu etal. (2013) simulated permeability as a spatially
correlated lognormally distributed random variable and studied its influence on water flow
and slope stability. Le etal. (2013) investigated rainfall-induced differential settlement of a
strip foundation on an unsaturated soil with spatially varying values of either preconsolida-
tion stress or porosity.
In this chapter, the basic theory and concepts of random field are presented in Section 8.2.
Some common methods of random field generation, such as matrix decomposition method,
fast Fourier transformation method, turning bands method (TBM), and so forth are reviewed
in Section 8.3. An example of stochastic modeling for seepage and stability of a two-dimen-
sional random heterogonous soil slope is presented in Section 8.4.

8.2 RANDOM FIELD THEORY

8.2.1 Concept of random field


A random field is a conceivable model to characterize continuous spatial fluctuations of a
soil property within a soil unit. The actual value of a soil property at each location within
the unit is assumed to be a realization of a random variable. Field can be modeled as a
Probabilistic assessment of randomly heterogeneous soil slopes 301

location-dependent random variable function Z(x), where x is the location coordinates


vector (x can be along any direction and is not necessarily along x-axis). The covariance/
autocovariance between two locations x and x is

C ( x, x ) = E {Z(x)Z(x)} E [ Z(x)] E [ Z(x)] (8.1)

where:
E[.] is the expectation operator

If the random field is stationary, the covariance of the soil property values at two locations
is only a function of the separation distance

C() = C ( x, x + ) = E {Z(x)Z(x + )} E [ Z(x)] (8.2)


2

Downloaded by [Monash University Library] at 23:45 02 August 2017

where:
is the lag vector representing separation between two spatial locations

The normalized form of the covariance function is known as the autocorrelation function (ACF):

C ()
() = (8.3)
C (0 )

where:
C(0) is the autocovariance function at zero separation distance, which is the stationary
variance of data

The above autocovariance function and ACF are usually estimated from soil samples from
a site. Analytical models are fitted to the sample ACFs using regression analysis. One of the
commonly used autocorrelation models is the exponential model, in which the autocorrela-
tion decays exponentially with the separation distance. The one-dimensional form of the
exponential model is

2
( ) exp (8.4)

where:
t is the separation distance between two points and d is the scale of fluctuation

The frequently used autocorrelation models are presented in Table8.1. The scale of fluctua-
tion d is an important indicator of soil spatial variability describing the distance within which
the spatially random values will tend to be significantly correlated (i.e., by more than about
10% [Fenton and Griffiths, 2008]). That is to say, within a range of d, soil property values are
highly correlated; but in a range over d, soil property values are irrelevant. Usually, the scale
of fluctuation is estimated using the moving window method or by curve fitting of empirical
autocorrelation models.

8.2.2 Spatial-averaged soil properties


Parameters in a geotechnical analysis usually refer to averages of soil properties over
some area or some volume; for example, average shear strength along a sliding surface
or average stiffness of a volume affected by loading. Hence, geotechnical analysis con-
cerns mostly the uncertainties in the averaged properties over specific surfaces or volumes.
302 Rainfall-Induced Soil Slope Failure

Table8.1 Common autocorrelation functions and variance reduction functions


Type Autocorrelation function Variance reduction function
2a 2 L L/ a
Exponential ( ) = e /a, a = 2 2 (L ) = a 1 + e
L2
b2 L L2 / b 2
( ) = e ( ) , b = 2 (L ) = b erf ( L b ) 1 + e
/ b 2
Gaussian
L2
1 1.5( /d ) + 0.5( /d ) ( d )
3
( ) = 2 ( L ) = 2 1.5 L d + 0.25 ( L d )
3
Spherical
0 ( > d )

1 1 L
Uniform type I ( ) = 2 (L ) =
0 > /L L>
Downloaded by [Monash University Library] at 23:45 02 August 2017

1 L /2
1 2
Uniform type II ( ) = 2 (L ) =
0 > 2 L (1 4L ) L > /2

1 L / ( 3 ) L
1
Triangular ( ) = 2 (L ) =
0 > (1 3L ) L>
L
Notes: d is scale of fluctuation; erf() is the error function; a, b, and d are coefficients of correlation function; L is the domain
size for the spatial-averaged soil property.

Thepoint-to-point variation forms the basis for quantitative assessment of uncertainties


of averaged soil parameters. Assume the variance of a soil property is varP. The larger the
domain size over which the property is averaged, the larger the reduction in the variance of
the spatial averaged soil property will be. The variance of the spatial averaged soil property,
varA , can be expressed as the product of the variance reduction factor function and the point
variance varP (Vanmarcke, 1977):

varA = 2 ()varP (8.5)


where:
G2(t) is the function of variance reduction factor, which can be determined based on
the ACF

The spatial averaged coefficient of variation (COV), covA , over a domain can be expressed as

cov A = ()cov P (8.6)

where:
covP is the point COV of a soil property

The variance reduction functions for the common ACFs are shown in Table8.1.

8.2.3 Definitions in geostatistics


In geostatistics, the spatial variability of soil properties are described by variogram or semi-
variogram function (Deutsch and Journel, 1997). The variogram is defined as the variance of
the difference {Z(x + ) Z(x)} .

2() = var {Z(x + ) Z(x)} (8.7)


Probabilistic assessment of randomly heterogeneous soil slopes 303

where:
var[] is the variance operator

The semivariogram function is defined as half of the variogram and can be expressed as the
difference of the stationary variance and covariance:

() = C(0) C() (8.8)

By definition, g(t) is semivariogram and the variogram is 2g(t). However, the terms semivar-
iogram and variogram are often used interchangeably.
The relationship between the variogram g(t) and covariance C(t) is shown in Figure8.3.
In general, g(t) increases from its initial value at g(0) as t increases. For continuous random
variables, the variogram function levels off and becomes stable about a limiting value called
the sill, which is generally at, or near, the variance of the stochastic process. The separation
Downloaded by [Monash University Library] at 23:45 02 August 2017

distance where the variogram function approaches the sill is called the range of influence.
In some cases, the behavior of the variogram near the origin (t = 0) can have a discontinu-
ity. The discontinuity near the origin is called the nugget effect, and produces an apparent
intercept at zero separation distance termed the nugget. The nugget effect can result from
small-scale effects or measurement errors.
Based on the variogram function or the ACF, the autocovariance distance, sometimes
named as correlation length, autocorrelation length, or correlation distance, is defined as
the distance at which the spatial variance/autocorrelation has decayed by 1/e (37%). The
scale of fluctuation is usually between 1.4 and 2.0 times the correlation length for expo-
nential, Gaussian, and spherical ACFs (Vanmarcke, 1983). In addition, the ACF is named
correlogram in geostatistics.

8.2.4 Reported statistics of spatial variability


Table8.2 summarizes reported values of scale of fluctuation for various soil properties from
literature. As shown in the table, for shear-strength-related parameters such as cone tip
resistance by Cone Penetration Test (CPT), Standard Penetration Test (SPT) N-value, und-
rained shear strength S u, the vertical scale of fluctuation is usually less than 5 m; however,

Sill
C(0) = 2

C()
()

Range of influence

()
Nugget

Separation distance,

Relationship between autocovariance function C() and variogram ().


Figure8.3
304 Rainfall-Induced Soil Slope Failure

Table8.2 Scale of fluctuation for various soil properties


Soil property Scale of fluctuation d Soil Source Test
Average cone h = 3560 m Marine clay Tang (1979) CPT
resistance (different levels)
Average cone h = 55 m (03 m
resistance below sea bottom)
CPT tip resistance, qc v = 2 m Sensitive clay Chiasson etal. (1995) CPT
CPT tip resistance, qc h = 20.035.0 m Glacial sands Vrouwenvelder and CPT
Calle (2003)
CPT tip resistance, qc v = 0.81.8 m Popescu etal. (1995)
CPT tip resistance, qc v = 1 m Clay Vanmarcke (1977)
CPT tip resistance, qc h = 1438 m Offshore soils Keaveny etal. (1989) CPT
Downloaded by [Monash University Library] at 23:45 02 August 2017

CPT tip resistance, qc h = 512 m Silty clay Lacasse and CPT


v = 1 m deLamballerie (1995)
CPT tip resistance, qc v = 3 m Clean sand Alonzo and Krizek CPT
v = 1 m Mexico clay (1975)
CPT tip resistance, qc v = 0.5 m Copper tailings Baecher and CPT
Christian(2003)
CPT tip resistance, qc v = 1.6 m Clean sand Kulatilake and CPT
Ghosh(1988)
CPT tip resistance, qc v = 0.12.2 m Sand, clay Phoon and Kulhawy CPT
h = 3.080.0 m (1996)
Corrected CPT tip v = 0.20.5 m Clay Phoon and Kulhawy CPT
resistance, qT h = 23.066.0 m (1996)
Penetrometer h = 40.070.0 m Sandy clay Mulla (1988)
resistance
DMT P0 h = 0.52.5 m Sand Jaksa etal. (2004) DMT
DMT P0 v = 1 m Varved clay DeGroot (1996) DMT
Undrained shear v = 2 m Chiasson etal. (1995) FVT
strength Su
Undrained shear v = 2.56 m Clay Asaoka and A-Grivas FVT
strength (1982)
Undrained shear h = 23 m Sensitive clay DeGroot and FVT
strength Su Baecher (1993)
Undrained shear v = 1 m Sensitive clay Baecher (1982) FVT
strength Su
Undrained shear v = 0.5 m Chicago clay Wu (1974) Unconfined
strength Su compression
test
Undrained shear v = 0.30.6 m Offshore soils Keaveny etal. (1989) Triaxial tests
strength Su and DSS tests
Undrained shear v = 0.86.1 m Clay Phoon and Kulhawy Laboratory
strength, Su (1996) test
Undrained shear v = 26.2 m Clay Phoon and Kulhawy VST
strength, Su h = 4660 m (1996)
(Continued)
Probabilistic assessment of randomly heterogeneous soil slopes 305

Table 8.2 (Continued) Scale of fluctuation for various soil properties


Soil property Scale of fluctuation d Soil Source Test
Unconfined v = 4 m, Honjo and Kuroda
compressive strength h=80.0m (1991)
Shear strength v = 2.0 m Clay Ronold (1990)
Shear strength v = 2.0 m, Clay Souli etal. (1990)
h=20.0m
N-value v = 2.4 m Sand Phoon and Kulhawy SPT
(1996)
N-value h = 20 m Dune sand Hilldale-Cunningham SPT
(1971)
N-value h = 17 m Alluvial sand DeGroot (1996) SPT
Permeability h = 1216 m Unlu etal. (1990)
Downloaded by [Monash University Library] at 23:45 02 August 2017

ln(kunsaturated)
ln(k) v = 3.2 m, Rehfeldt etal. (1992) Flowmeter
h=25.0m
ln(k) v = 1.53.0 m, Rehfeldt etal. (1992)
h=25.050.0 m
ln(k) v = 0.21 m, Hess etal. (1992) Large-scale
h=210 m tracer test
k h = 1500 m Salt dome Ditmars etal. (1988)
k h = 0.52 m Compacted clay Benson (1991)
k h = 12.5 m Sand Bjerg etal. (1992)
Natural water v = 1.612.7 m Clay, loam Phoon and Kulhawy
content, wn (1996)
Natural water h = 170.0 m Clay Phoon and Kulhawy
content, wn (1996)
Liquid limit, wL v = 1.68.7 m Clay, loam Phoon and Kulhawy
(1996)
Total unit weight,g v = 2.47.9 m Clay, loam Phoon and Kulhawy
(1996)
Effective unit weight, v = 1.6 m Clay Phoon and Kulhawy
g (1996)
Sand content h = 60.080.0 m Sandy clay Mulla (1988)
Clay content h = 40.060.0 m Sandy clay Mulla (1988)
Natural water h = 40.060.0 m Sandy clay Mulla (1988)
content
Thickness of natural h = 750 m Rosenbaum (1987)
deposit
Notes: CPT, Cone Penetration Test; SPT, Standard Penetration Test; DMT, Dilatometer Test; FVT, Field Vane Test; DSS, Direct
Simple Shear test.

the horizontal one is generally greater than 10 m. For basic soil properties such as water
content and unit weight, the vertical scale of fluctuation is slightly larger but basically is less
than 10 m. The horizontal scale of fluctuation can be greater than 100 m. For permeability
of a soil, the scale of fluctuation for ln(k) in vertical direction can be less than 1 m and the
horizontal scale of fluctuation is usually greater than 10 m.
306 Rainfall-Induced Soil Slope Failure

8.3 MODELING OF RANDOM FIELD

Different algorithms are available for simulation of random field. The most popular meth-
ods include the spectral methods based on Fourier transformation (Cooley and Tukey,
1965; Mejia and Rodriguez-Iturbe, 1974; Dietrich and Newsam, 1996), the moving average
method (Matrn, 1986; Yaglom, 1987; Oliver, 1995; Chiles and Delfiner, 1999; Cressie and
Pavlicov, 2002), the local average subdivision (LAS) method (Fenton and Vanmarcke, 1990;
Fenton and Griffths, 2008), the covariance matrix decomposition method (Davis, 1987),
the turning bands method (TBM) (Matheron, 1973; Mantoglou and Wilson, 1982), and the
sequential simulation methods (Bellin and Rubin, 1996; Deutsch and Journel, 1997). If prop-
erty values are known or measured at specific points, the conditional random field simulation
methods (Davis, 1987; Gutjahr etal., 1997; Strebelle, 2002; Emery, 2008) can be applied.
Comprehensive detailed review on the various methods can be found in Stefanou (2009),
Deutsch and Journel (1997) and Fenton and Griffiths (2008), etc.
Downloaded by [Monash University Library] at 23:45 02 August 2017

8.3.1 Covariance matrix decomposition method


The covariance matrix decomposition method is a direct method to produce a random field
assuming that the parameters at different locations in the field are correlated random vari-
ables. This method is applicable to any configuration of the simulated locations and any
covariance model.
For simplicity, consider a one-dimensional problem. The one-dimensional space has first
been discretized into n points with the soil property at each point is defined as a random vari-
able. Assume that the prescribed covariance matrix for these correlated random variables is
C. If C is positive definite, then a correlated standard normal random field Z can be produced
using independent standard normal random variables according to the following equation:

Z = LU (8.9)
where:
L is a lower triangular matrix satisfying LLT = C
U is a vector of n-independent standard normal random variables
Z is a vector of n-correlated standard normal random variables

The lower triangular matrix L is typically obtained using the Cholesky decomposition
method. If a soil property Z of the random field is normally distributed with a known mean
and variance, then a realization for the soil property can be obtained using the realization
of the correlated standard normal variate Z as follows:

Z = ZI + ZZ (8.10)

where:
I is the unit matrix
mZ and sZ are the mean value and standard variation of the normally distributed Z

Non-normal random fields can be obtained through a suitable transformation of a nor-


mally distributed random field. For example, a lognormally distributed random field can be
obtained from

Z = exp ( InZI + InZZ ) (8.11)


Probabilistic assessment of randomly heterogeneous soil slopes 307

where InZ and ln Z are the mean and standard variation of ln(Z), respectively.
Compared with the TBM, spectral fast Fourier method and the sequential Gaussian simu-
lation (SGS) method, this method shows the least artificial bias (Harter 1994). It is usually
used for small fields with small size of discrete points as it requires the decomposition of
a covariance matrix. For random fields with many realizations, this is indeed a very effec-
tive way of random field generation, because the covariance matrix must only be decom-
posed once for an entire simulation (Harter, 1994). Each realization then simply requires
the generation of random numbers and the multiplication with the L matrix (Equation 8.9).
It should yet be noted that considerable round-off error may be induced in the Cholesky
decomposition when covariance matrices are poorly conditioned and become numerically
singular.

8.3.2 Fourier transformation methods


Downloaded by [Monash University Library] at 23:45 02 August 2017

The Fourier transform method are based on the spectral representation of continuous ran-
dom field, Z(x), which can be expressed as follows (Yaglom, 1962):


Z(x) =

eixdW () (8.12)

where:
Z is the soil property
x is the vector of spatial variable
w is the angular frequency
dW() is an interval white noise process with mean zero and variance of S(w)dw

S(w) is the spectral density function of the random field. Z(x) and is simply the Fourier
transform of the covariance of Z(x) (Harter, 1994). The term S(w)dw is a measure of the
contribution of the amplitude of a frequency w to the random field.
The above integral is usually expressed as a summation. For illustration purpose, only one-
dimensional case will be presented. Based on the discrete Fourier transform (DFT) method, the
random field in one dimension that is discretized into 2K+1 frequencies can be expressed as
K


Z(x) = A cos( x) + B sin( x) (8.13)
k= K
k k k k

where:
x is the location coordinate
k is the kth discrete angular frequency
Ak and Bk are mean zero, mutually independent, normal distributed random numbers

The variance of Ak and Bk can be obtained from the spectral density function S(w). Hence, the
spectral representation of a single random field realization can be intuitively understood as a
field of amplitudes, where the coordinates are the frequencies of sinecosine waves.
In the fast Fourier transform (FFT) method (Cooley and Tukey, 1965), both space and fre-
quency are discretized into a series of points. Assume the numbers of spatial and frequency
discretization points are the same and the random field is periodic (i.e., Z(xj) = Z(xK+j). For
discrete Z(xj), j = 1, 2, , K, the Fourier transform can be evaluated as
K


Z(x j ) = e
k=0
k
i (2 jk / K )
(8.14)
308 Rainfall-Induced Soil Slope Failure

where k are the Fourier coefficients with k = Ak iBk . Ak and Bk are mean zero, mutu-
ally independent, normal distributed random numbers, whose variance can be obtained
from the spectral density function S(w). Ak and Bk are symmetric due to the fact that Z
is real.
The simulation process of DFT and FFT methods is as follows.

1. Decide how to discretize the space and frequency of spectral density function.
2. Generate independent normally distributed realizations of Ak and Bk.
3. Use the symmetry relationships to produce the Fourier coefficients (only in FFT).
4. Produce the field realization using Equation 8.13 or 8.14.

The major shortcoming of DFT is its computational demand. The DFT approach is basically
only computational practical for one-dimensional problem (Fenton and Griffiths, 2008).
Downloaded by [Monash University Library] at 23:45 02 August 2017

The FFT is generally much faster than DFT and can be adopted for two-dimensional or
three-dimensional problems. However, the covariance function obtained from the FFT ran-
dom field simulation is always symmetric about the midpoint of the field. This deficiency can
be overcome by generating a field twice as long as required in each coordinate direction and
keeping only the first quadrant of the field (Fenton and Griffiths, 2008).

8.3.3 Turning bands method


The TBM is a multidimensional random field generation method that performs simula-
tions along one-dimensional lines (bands) instead of synthesizing the multidimensional field
directly. The TBM generally includes the following steps (Matheron, 1973; Mantolou and
Wilson, 1982; Fenton and Griffiths, 2008):

1. Define the multidimensional covariance model C() of the random field in which is
the distance vector between the two points in space.
2. Determine the one-dimensional covariance C1() of the random field. Mantoglou and
Wilson (1982) and Bras and Rodriyguez-Iturbe (1985) review and summarize several
multidimensional covariance models and the corresponding unidimensional covari-
ance functions.
3. Choose an arbitrary origin within or near the domain of the field to be generated.
Select L lines crossing the domain with the ith line having a direction given by the unit
vector ui.
4. Generate a realization of one-dimensional random field, Zi(), along the ith line base on
the one-dimensional covariance function C1(), where is the coordinate along the line.
5. Project one field point xk orthogonally onto each line to define the coordinate ki
(ki = x k ui ) and obtain the component of the one-dimensional random field realiza-
tion value Zi (ki ) = Zi (x k ui ).
6. Summing the contributions of all lines and normalize to the field value Z(x k ) by the
factor L, that is, Z(x k ) = 1 L iL=1 Zi (x k ui ) for each xk.

The TBM method is not a fundamental random field generator and requires an existing
one-dimensional generator such as FFT. Therefore, the TBM method can be classified based
on different approaches of generation one-dimensional line processes: (1) the approach that
uses a moving average method combined with weighting function approximation (Bras
et al., 1985), (2) the approach that evaluates the integration of unidimensional spectral
density function by a standard integration approximation method (Mantolou and Wilson,
Probabilistic assessment of randomly heterogeneous soil slopes 309

1982; Vanmarcke, 1983), and (3) the approach that applies an FTT algorithm to generate
the unidimensional line processes (Tompson etal., 1989; Kottegoda and Kassim, 1991).

8.3.4KarhunenLoeve expansion method


The KarhunenLoeve (KL) expansion of a random field (Ghanem and Spanos, 1991; Huang
etal., 2001; Phoon etal., 2002a,b; Yang etal., 2004) is based on the spectral decomposition
of the covariance function:


Z(x, ) = (x) +
i =1
i i (x)i () (8.15)
Downloaded by [Monash University Library] at 23:45 02 August 2017

where:
x is the coordinate of a point in the continuous domain
is the coordinate in the sample space
( x ) is the mean function of the random field
i(q) is uncorrelated standard normal random variables
ji and li denote the eigenfunctions and eigenvalues of the covariance matrix obtained
from solving the homogeneous Fredholm integral equation of the second kind:


C(x, x) (x)dx = (x) (8.16)

i i i

The approximate random field is defined by truncating the ordered series:

M
Z(x, ) = (x) +
i =1
i i (x)i () (8.17)

where M is the truncating level.


The truncated KL expansion underestimates the true variability of the original random field.
The truncating level M to be chosen strongly depends on the desired accuracy and the covari-
ance function of the random field. In addition, the truncated KL expansion of homogeneous
random fields is only approximately homogeneous, because the standard deviation function
of the truncated field varies in space (Stefanou and Papadrakakis, 2007; Betz etal., 2014).

8.3.5 LAS method


The LAS method proposed by Fenton and Vanmarcke (1990) is a fast and accurate method
of producing realizations of a random process. The motivation of the method arose out
of the fact that most engineering measurements are only defined over some finite domain
and thus represent a local average of the property. The method was based on the stochastic
subdivision methods (Carpenter, 1980; Fournier etal., 1982) and incorporates the concept
of local averaging. The procedure of LAS for a one-dimensional stationary random field is
presented here to illustrate the basic concept of this method.
The construction of a local average process is essentially a topdown recursive fash-
ion (Figure8.4). In Stage 0, a global average is generated for the process. In Stage 1, the
domain is subdivided into two regions whose local averages must in turn average to the
global (parent) value. Subsequent stages are obtained by subdividing each parent cell and
310 Rainfall-Induced Soil Slope Failure

Stage 0 Z 10

Stage 1 Z 11 Z 21

Stage 2 Z 12 Z 22 Z 32 Z 42

Stage 3 Z 13 Z 23 Z 33 Z 43 Z 53 Z 63 Z 73 Z 83

Stage 4

Top-down approach of the local average subdivision (LAS) method. (Adopted from Fenton, G. A.,
Figure8.4
and Vanmarcke, E. H., Journal of Engineering Mechanics, 116, 17331949, 1990. With permission.)

generating values for the resulting two regions while preserving upwards averaging. The
Downloaded by [Monash University Library] at 23:45 02 August 2017

global average remains constant throughout the subdivision. The detailed procedures are
as follows.

1. Generate a normally distributed global average Z10 with mean zero and variance
obtained from local averaging theory.
2. Subdivide the field into two equal parts and generate two normally distributed values
of Stage 1, Z11 and Z21 , whose means and variances satisfy the following three criteria:
a. Z11 and Z21 show the correct variance according to local averaging theory.
b. Z11 and Z21 average to the parent value, that is, 1 2 ( Z11 + Z21 ) = Z10. In other words,
the distributions of Z11 and Z21 are conditioned on the value of Z10.
c. Z11 and Z21 are properly correlated.
3. Subdivide each cell in Stage 1 into equal parts. Generate normally distributed values,
Z12, Z22, Z32, and Z42 of Stage 2. The means and variances of Z12, Z22, Z32, and Z42 satisfy the
above three criteria just like in Stage 1. In addition, Z12 and Z22 are properly correlated
with Z32 and Z42, which can be satisfied approximately by conditioning the distributions
of Z12 and Z22 also on Z21 .

8.3.6 Sequential simulation method


Different sequential simulation methods use basically the same algorithm and can be clas-
sified based on different data types. Sequential Gaussian simulation (SGS) is a simulation
method for continuous variables. Sequential indicator simulation (SIS) simulates discrete
variables, using SGS methodology to create a grid of zeros and ones. The general sequential
simulation procedure is as follows (Deutsch and Journel, 1997).

1. Perform a normal transformation of the raw data into data with standard normal
cumulative density function (CDF).
2. Define a random path that visits each node of the grid (not necessary regular) once.
3. Select one node that is not yet simulated in the grid. Construct the conditional prob-
ability distribution function by kriging.
4. Draw a simulated value from the conditional probability distribution function.
5. Include the newly simulated value in the data set and check whether the results honor
the data and the desired spatial variability.
6. Repeat until all grid nodes have simulated values.
Probabilistic assessment of randomly heterogeneous soil slopes 311

8.4SEEPAGE AND STABILITY OF A RANDOMLY HETEROGENEOUS


SLOPE UNDER RAINFALL INFILTRATION

The permeability of unsaturated soil may change spatially due to uncertainties in soil fabric.
Itis of great significance to perform probabilistic infiltration and stability analysis of slopes
considering the permeability function as a random field. Few attempts have been made to
study the variations in groundwater table in a spatial varied soil slope under rainfall infil-
tration. The statistical response of pore-water pressures in two-dimensional unsaturated
heterogeneous slopes has not been sufficiently investigated. Also the critical hydraulic con-
ditions that may lead to failure of heterogeneous slopes are not well known. The objective
of this section is to illustrate how the variability of permeability function propagates to the
variability of hydraulic conditions (i.e., pore-water pressure and groundwater table) induced
by steady rainfall infiltration and its effect on slope stability.
Downloaded by [Monash University Library] at 23:45 02 August 2017

8.4.1 Slope geometry and boundary conditions


Numerical modeling of seepage and slope stability of a hypothetical slope is conducted.
Figure8.5 shows the geometry of the slope. The hydraulic heads on both side boundaries
which are below the groundwater table (BC and DE) are constant. The bottom boundaries
as well as the side boundaries above the water table (AB and EF) are assumed to be imper-
meable. The uncertainties of boundary conditions are not considered.

8.4.2 Soil properties and generation of random field


The Fredlund and Xing (1994) SWCC model is used in this example. The exponential equa-
tion proposed by Leong and Rahardjo (1997) is employed for the unsaturated permeability
function,
lr
r
k = ks w (8.18)
s r

q
60 X
A H Y
50
H = 10 m
Elevation (m)

40 30 G
B
F
30
E
20
C D
10 X Y

0
20 10 0 10 20 30 40 50 60 70 80 90 100 110 120
Horizontal distance (m)

Geometry of a hypothetic slope in the numerical study.


Figure8.5
312 Rainfall-Induced Soil Slope Failure

where:
ks is the saturated permeability
qw is the volumetric water content
qs is the saturated volumetric water content
qr is the residual water content
(qw qr)/(qs qr) is the normalized water content
dlr is a constant depending on the soil type

The SWCC and unsaturated permeability function are presented in Figure8.6 with the soil
parameters listed in Table8.3. The soil within the slope is assumed to be a fine sand with
the mean value of ks is equal to 2 105 m/s. The lr value in the permeability function is
assumed to be 3 (Tami etal., 2004). The vertical flux applied to the slope surface is givenas
2 107 m/s (moderate rainfall intensity), that is, the ratio between the rainfall intensity and
Downloaded by [Monash University Library] at 23:45 02 August 2017

the saturated permeability is 0.01. This small value is shown to be able to maintain a con-
stant matric suction within the unsaturated zone (Zhang etal., 2004).
As the most important parameter in rain infiltration in unsaturated soil slope is the satu-
rated permeability, the spatial variability of ks is considered and the other soil hydraulic
parameters remain constant for different locations within the slope. The shear strength
parameters are also considered to be deterministic.
The generation of a random field normally requires a mean value (m), a standard devia-
tion or COV and a correlation structure. Based on field measurements, previous studies have

0.4
Volumetric water content

0.3

0.2

0.1 Residual water


content
0.0
1.0E01 1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04
(a) Soil suction (kPa)
1.0E04

1.0E05
Permeability (m/s)

1.0E06

1.0E07

1.0E08

1.0E09

1.0E10
1.0E01 1.0E+00 1.0E+01 1.0E+02 1.0E+03 1.0E+04
(b) Soil suction (kPa)

Soil property functions used in the numerical model: (a) soilwater characteristic curve and
Figure8.6
(b)permeability function. (From Zhu, H. et al., Computers and Geotechnics, 48, 249259, 2013.
With permission.)
Probabilistic assessment of randomly heterogeneous soil slopes 313

Table8.3 Parameters for the hypothetic study cases


Parameters Definition Value
ks Saturated coefficient of permeability k = 2 105 m/s
S

COV = 60%100%
Correlation length (d): 0.5, 8, 100, 1000, 2000 m
Normalized correlation length = d/H: 0.025, 0.4,
5, 10, 100 m
qs Saturated volumetric water content 0.4
qr Residual volumetric water content 0.02
af SWCC parameter 5
nf SWCC parameter 2
mf SWCC parameter 1
dlr Permeability function parameter 3
Downloaded by [Monash University Library] at 23:45 02 August 2017

q Vertical infiltration flux 2 10 7 m/s


c' Effective cohesion 1 kPa
f' Effective angle of internal friction 30
fb Angle indicating the rate of increase of 15
shear strength related to soil suction

shown that ks can be modeled as a lognormal random field (Whitman, 2000), with a mean
of ks and a standard deviation of ks . Thus, ln ks is assumed to follow a normal distribution
with a mean of lnks and a variance of, 2 lnks where

= ln[1 + ( ks ) / ( ks ) ] (8.19)
2 2
2 ln ks

1 2
ln ks = ln(ks ) ln ks
2 (8.20)

In Equations 8.19 and 8.20, 2 lnks is dimensionless. In this study, an isotropic spherical auto-
correlation model is assumed.
A generic range of COV of ks has been suggested as 60%100% in the literature (Duncan,
2000), and the correlation length of ln ks is often taken as 0.15 times the slope height
(Srivastava et al., 2010). In order to comprehensively investigate the perfect correlation
scenario where correlation length is sufficiently high to yield a nearly homogeneous domain,
the correlation length is extended to a range of 0.52000 m or 0.025100 times the slope
height, as shown in Table8.3. To avoid the effect of slope dimension, a normalized correla-
tion length (m) is used, which is defined as m = d/H (H is the slope height; d is the coef-
ficient of the spherical correlation model that represents the correlation length as shown in
Table 8.1).
In order to create a random field of a soil parameter, a rectangular random field is first
generated. The FFT code modified from Kozintsev (1999) is adopted. A sequence of data,
representing a realization of the Gaussian random field, is first generated with parameters
represented by zero mean, unit variance, and a spherical correlation structure. A lognor-
mally distributed random field of ks is then transformed from the Gaussian random field.
Figure8.7 presents a series of realizations for the Gaussian field, the random fields of ln ks
and ks. The random fields of the Gaussian field and ln ks are linearly related. The random
field of ks is truncated to the geometry of the slope for numerical modeling as shown in
Figure8.8.
314 Rainfall-Induced Soil Slope Failure

40

35 2

30 1

25
0
20
y
1
15
2
10

5 3

10 20 30 40 50 60 70 80 90 100
Downloaded by [Monash University Library] at 23:45 02 August 2017

(a) x
55 9.0

50 9.5
10.0
45
10.5
40 11.0
y (m)

35 11.5
12.0
30
12.5
25 13.0
20 13.5
14.0
15
0 20 40 60 80 100
(b) x (m)
105
55

50 14

45 12

40 10
y (m)

35 8

30 6

25 4

20 2

15
0 20 40 60 80 100
(c) x (m)

Realizations of random field (k = 2 105 m/s, d = 8 m, COV = 100%): (a) Gaussian field,
Figure8.7 S

(b) log-permeability random field, and (c) permeability random field. (From Zhu, H. et al.,
Computers and Geotechnics, 48, 249259, 2013. With permission.)
Probabilistic assessment of randomly heterogeneous soil slopes 315

Elevation (m)

Horizontal distance (m)


Downloaded by [Monash University Library] at 23:45 02 August 2017

Figure8.8Schematic of mapping the k s random field to a slope profile. (From Zhu, H. et al., Computers and
Geotechnics, 48, 249259, 2013. With permission.)

8.4.3 Methodology of stochastic modeling


8.4.3.1 Seepage and slope stability analysis
The PDE based finite element modeling software FlexPDE (PDE solutions Inc., 2015) is uti-
lized to solve the governing equation of infiltration. The software allows users to input vari-
ables in tabulated files by which stochastic finite element analyses are performed repeatedly.
The slope stability analysis is conducted using the program SVSlope (SoilVision System Ltd.,
2010) in which the pore-water pressure distribution are imported from seepage analysis. The
Bishops simplified method with the extended MohrCoulomb criterion for unsaturated soil
by Fredlund etal. (1978) is used in the slope stability analysis. The shape of the slip surface
is assumed to be circular. The critical slip surface corresponding to the minimum factor of
safety is searched in each slope stability analysis.

8.4.3.2Mapping of saturated permeability on


finite element seepage analysis
The ks value at each grid point of the rectangular random field grid is transferred into the finite
element seepage analysis as an input soil parameter. Bilinear interpolation is performed to map
from the rectangular random field grid to the finite element mesh.
Figure 8.9 shows that the range of ks within a slope varies with the correlation length.
Taking an arbitrary realization for example, the values of ks range from 0.02 104 m/s to
2.42 104 m/s at a correlation length of 0.5 m, and from 1.01 106 m/s to 3.53 106 m/s at
a correlation length of 1000 m. A larger correlation length leads to a smaller spatial differ-
ence in the values of ks.

8.4.3.3 Stochastic analysis by Monte Carlo simulation technique


In this example, the Monte Carlo simulation (MCS) technique is adopted for stochastic anal-
ysis. Realizations of random field are first generated by the FFT random field generation
method. Then the deterministic problem is solved for each realization of the random field
316 Rainfall-Induced Soil Slope Failure

max 2.42
k: 2.1
j: 1.9
i: 1.7
Elevation (m)
h: 1.5
g: 1.3
40 m f:
e:
1.1
0.9
d: 0.7
c: 0.5
b: 0.3
100 m a: 0.1
min 0.02
Scale 104
(a) Horizontal distance (m)
max 6.25
k: 6.0
j: 5.1
i: 4.5
h: 3.6
Elevation (m)

g: 3.3
Downloaded by [Monash University Library] at 23:45 02 August 2017

f: 3.0
40 m

e: 2.7
d: 2.4
c: 1.5
b: 1.2
a: 0.9
100 m min 0.69
Scale 105
(b) Horizontal distance (m)
max 3.53
k: 3.4
j: 3.3
i: 3.2
h: 3.1
Elevation (m)

g: 3.0
f: 2.6
40 m

e: 2.3
d: 2.1
c: 1.3
b: 1.2
a: 1.1
100 m min 1.01
Scale 106
(c) Horizontal distance (m)

Figure8.9Typical realizations of random fields with various correlation lengths ( ks = 2 10 5m/s, COV =
100%): (a) d = 0.5 m, (b) d = 100 m, and (c) d = 1000 m. (From Zhu, H. et al., Computers and
Geotechnics, 48, 249259, 2013. With permission.)

and a population of the random response quantities is obtained. This population can then be
used to obtain statistics of the prediction. In this example, the number of realizations by the
MC simulation is 1000.
A parametric analysis is carried out to study the sensitivity of pore-water pressure and
factor of safety to the correlation length of ln ks. A base case (i.e., a homogeneous profile
with the mean value of ks) is analyzed for comparison purposes. In the graphs of results in
the following section, the results of the base case is denoted as deterministic.

8.4.4 Results and discussion


8.4.4.1 Influence of soil spatial variability on pressure head profiles
Figures 8.10 and 8.11 show the estimated quantiles of the pore-water pressure head
profiles along cross section XX at the crest and section YY at the middle of the slope
(Figure8.5), respectively. It can be seen that a higher correlation length produces a smaller
25% quantile of the pressure head profiles, but larger 50% and 75% quantiles. The quantiles
of pressure head at normalized correlation lengths of 50 and 100 m are almost the same.
Probabilistic assessment of randomly heterogeneous soil slopes 317

65

55

Elevation (m)
45

35 m = 0.025
m = 0.4
m=5
25 m = 50
m = 100
Deterministic
15
8 6 4 2 0 2 4
(a) Q25% of pore-water pressure head (m)
Downloaded by [Monash University Library] at 23:45 02 August 2017

65

55
Elevation (m)

45

35 m = 0.025
m = 0.4
m=5
25 m = 50
m = 100
Deterministic
15
6 4 2 0 2 4
(b) Q50% of pore-water pressure head (m)
65

55
Elevation (m)

45

35 m = 0.025
m = 0.4
m = 0.5
25 m = 50
m = 100
Deterministic
15
6 4 2 0 2 4
(c) Q75% of pore-water pressure head (m)

P ressure head profiles along section XX at steady state for various correlation lengths of
Figure8.10
ln ks: (a) 25% quantile, (b) 50% quantile, and (c) 75% quantile. (From Zhu, H. et al., Computers
and Geotechnics, 48, 249259, 2013. With permission.)

Hence, a normalized correlation length of 100 represents a reasonable value that represents
perfect correlation of spatial variability. The depth within which a constant suction can
be maintained under steady-state infiltration increases with increasing of the correlation
length. Santoso etal. (2011) conducted a one-dimensional stochastic infiltration analysis for
an infinite slope assuming a constant groundwater table. They found that the pore-water
318 Rainfall-Induced Soil Slope Failure

55

45
Elevation (m)

35
m = 0.025
m = 0.4
25 m=5
m = 50
m = 100
Deterministic
15
6 4 2 0 2
Downloaded by [Monash University Library] at 23:45 02 August 2017

(a) Q25% of pore-water pressure head (m)


55

45
Elevation (m)

35
m = 0.025
m = 0.4
m=5
25 m = 50
m = 100
Deterministic
15
6 4 2 0 2
(b) Q50% of pore-water pressure head (m)
55

45
Elevation (m)

35

m = 0.025
m = 0.4
25 m=5
m = 50
m = 100
Deterministic
15
6 4 2 0 2
(c) Q75% of pore-water pressure head (m)

Figure 8.11Pressure head profiles along section YY at steady state for various correlation lengths of
lnks:(a) 25% quantile, (b) 50% quantile, and (c) 75% quantile. (From Zhu, H. et al., Computers
and Geotechnics, 48, 249259, 2013. With permission.)
Probabilistic assessment of randomly heterogeneous soil slopes 319

60
Pressure bound
Q25%, m = 0.025
Q75%, m = 0.025
Q25%, m = 0.4
50
Q75%, m = 0.4
Q25%, m = 50
Elevation (m)

Q75%, m = 50
40 Q25%, m = 100
Q75%, m = 100

30

20
Downloaded by [Monash University Library] at 23:45 02 August 2017

6 4 2 0 2 4 6
(a) Pore-water pressure head (m)
55

45
Elevation (m)

35

Q25%, m = 0.025
25 Q75%, m = 0.025 Q75%, m = 50
Q25%, m = 0.4 Q25%, m = 100
Q75%, m = 0.4 Q75%, m = 100
Q25%, m = 50 Deterministic
15
6 4 2 0 2
(b) Pore-water pressure head (m)

Figure8.12Bounds of pore-water pressure head along: (a) section XX, (b) section YY. (From Zhu, H.
etal., Computers and Geotechnics, 48, 249259, 2013. With permission.)

pressure profile changes with the statistical parameters of ks and finally reaches a perfect
correlation state when the correlation length is sufficiently long.
Another important indicator of variability in hydraulic conditions is the bound of pore-
water pressure head (Figure8.12), which represents the range of pore-water pressure head
in the stochastic simulation. The Q25% and Q75% bounds of the pressure head for cross
sections XX and YY are shown in Figure8.12. The spatial variability of ks can induce
a variation in pore pressure head of more than 3 m. A higher correlation length produces
more widely distributed negative pressures, which implies larger uncertainty. As shown in
Figure8.12(b), the matric suctions in the random heterogeneous slopes are within the range
of 50% to 125% of those in the deterministic slope.
320 Rainfall-Induced Soil Slope Failure

Width of pressure bound (m)


3

1 Crest
Slope

0
0 10 20 30 40 50
Normalized correlation length

Figure8.13Width of pressure bound versus normalized correlation length of ln ks.


Downloaded by [Monash University Library] at 23:45 02 August 2017

The relationships between the width of the pressure head bound and the correlation
length of ln ks are plotted in Figure8.13. The bound width increases sharply with increas-
ing normalized correlation length when m is less than 5. Relatively little change is observed
when the normalized correlation length is greater than 5. It can be noted that the pressure
bounds at the crest (cross section XX) are slightly wider than those at the middle of the
slope (crosssection YY).

8.4.4.2Influence of soil spatial variability on


variations of groundwater table
In addition to irregularity of pressure distribution in random heterogeneous soils, the
variation of groundwater table is another indicator of the performance of the slope. The
shape of the groundwater table in a heterogeneous soil differs from that in a homogeneous
soil. Figure8.14 shows two boundary envelopes that cover the possible locations of the
groundwater tables obtained from 1000 realizations for case with m = 50 and COV =
100%. The upper bound represents the highest position that the free water can reach, and
the lower bound is the lowest one. The variability of groundwater table is characterized

Upper bound
Lower bound
Mean
Elevation (m)

A
D
B

Horizontal distance (m)

Envelopes of the groundwater table for a case with m = 50, COV = 100%. (From Zhu, H. et al.,
Figure8.14
Computers and Geotechnics, 48, 249259, 2013. With permission.)
Probabilistic assessment of randomly heterogeneous soil slopes 321

14

Maximum groundwater table


12

10

difference D (m)
8

0
0 10 20 30 40 50
Normalized correlation length
Downloaded by [Monash University Library] at 23:45 02 August 2017

Variation of the maximum difference of groundwater table with the correlation length of ln k s.
Figure8.15

by the maximum difference (D) between these two curves in Figure8.14. A larger D value
implies a higher degree of variation.
Figure8.15 shows the relationship between the variability of the groundwater table and
the correlation length of ln ks. It is noted that D increases sharply with small values of
correlation lengths and changes slowly at high correlation length values. The D value is
very small at small correlation lengths, indicating that the groundwater table in this case
is similar to the deterministic case and the fluctuations can be ignored. This is because the
local averaging effect is significant at very small correlation lengths. Any point at which the
permeability is large will be surrounded by points where the permeability values are small.
As the correlation length increases, the groundwater table becomes wavy due to the presence
of a larger area of zones with similar soil property.

8.4.4.3 Effects of correlation length of ln k s on factor of safety


Figure 8.16(a) presents the variation of mean Fs with correlation length by analyzing 50,
100, 150, and 200 realizations (N = number of realizations). The trend is found to be insen-
sitive to the number of realizations and the mean Fs approaches the deterministic value at a
normalized correlation length of 0.025. Hence, 200 simulations are considered sufficient.
In Figure8.16(a), the factor of safety calculated in the deterministic case is higher than the
mean factors of safety considering the spatial variability of permeability function. This is in
good agreement with Griffiths and Fenton (2004) that probabilistic slope stability analysis
that ignores the spatial variability of soil properties underestimates the probability of the
slope failure. The greatest reduction in Fs occurs when d is of the same order of the slope
height, H. It is hypothesized that d 5H leads to the greatest reduction in Fs because this d
value allows enough spatial averaging along a failure surface.
In Figure8.16(b), the COV of factor of safety is sensitive to changes in the correlation
length of ln ks when the correlation length is small, and increases gently when correlation
lengths five times longer than the slope height. At a high level of correlation length, the area
within which the permeability functions are highly correlated covers the most part of the
slope profile. Therefore, the large variation of pore-water pressure distributions leads to a
large value of COV for the factor of safety.
The distribution of the factors of safety from 200 simulations is presented in Figure8.17(a).
The frequency histogram and the theoretically fitted lognormal probability densityfunction
322 Rainfall-Induced Soil Slope Failure

1.20 7

Deterministic Fs = 1.196 6
5

COV of Fs (%)
1.18
Mean of Fs

4
Number of realizations
N = 50 3
1.16
N = 100 2
N = 150
N = 200 1
1.14 0
0 10 20 30 40 50 0 10 20 30 40 50
(a) Normalized correlation length (b) Normalized correlation length

Statistical results of factor of safety with spatially variable saturated permeability parameter
Figure8.16
(COV = 100%): (a) variation of mean Fs with the correlation length of ln k s and (b) change in
Downloaded by [Monash University Library] at 23:45 02 August 2017

COV of Fs with the correlation length of ln ks.

1.4 0.20
Histogram
1.3
Probability density

0.16
Lognormal
Factor of safety

1.2 0.12 distribution


1.1
0.08
1.0 Deterministic Fs = 1.196
0.04
0.9
0.00
0.8

1. 05

1. 06
1. 1
1. 55
1. 78
1. 03
1. 29
1. 54
1. 79
1. 04

1. 56
1. 81

2
1. 0

1. 0
23
0 50 100 150 200 250

33
13

18
2

3
1
9
0
0
0
0
1

2
2
0.

(a) ith simulation (b) Factor of safety

Distributions of factor of safety at a correlation length of m = 50: (a) upper and lower bound of
Figure8.17
Fs and (b) frequency histogram of Fs. (Modified from Zhu, H. et al., Computers and Geotechnics,
48, 249259, 2013. With permission.)

ofFs are shown in Figure8.17(b). The data of the factor of safety are tested by the Kolmogorov
Smirnov (KS) goodness-of-fit test, in which the empirical cumulative frequency and the
corresponding theoretical CDF for the normal and lognormal distributions are calculated.
The maximum discrepancies of the two cumulative frequencies are Dmax = 0.079 for the
normal distribution and Dmax = 0.070 for the lognormal distribution. At the 5% significance
level, the critical value of D 0.05 is 0.096. As the maximum discrepancies for both normal and
lognormal distributions are less than 0.096, the two distribution types are both acceptable.
The lognormal distribution is better because the Dmax value is smaller.
Comparing the results from all the realizations, the values of Fs of the soil slope varies
from 0.909 to 1.357, which implies a range of 24% to +13.5% error in estimation of Fs
if the spatial variability is not considered for this specific slope (30 in slope angle, 20 m
in height and 2 105 m/s in mean saturated permeability). Previous studies (Tartakovsky
etal., 2003; Srivastava etal., 2010; Santoso etal., 2011) focus on the mean factor of safety
to illustrate the impact of soil spatial variability. The extreme values of Fs can demonstrate
the variability of slope safety more clearly. The results imply that design approaches using
mean values tend to be risky by ignoring the influence of spatially variable permeability
functions.
Probabilistic assessment of randomly heterogeneous soil slopes 323

max 3.53
k: 3.5
j: 3.3
i: 3.1
h: 2.9
g: 2.7
f: 2.5
e: 2.2
d: 2.0
c: 1.7
b: 1.4
a: 1.1
min 1.01
Scale 106
(a)
max 1.85
k: 1.8
j: 1.5
i: 1.4
h: 1.3
Downloaded by [Monash University Library] at 23:45 02 August 2017

g: 1.2
f: 1.1
e: 1.0
d: 0.9
c: 0.8
b: 0.7
a: 0.6
min 0.61
Scale 104
(b)

Realizations of ks and critical slip surfaces corresponding to: (a) minimum factor of safetyand
Figure8.18
(b) maximum factor of safety (m = 50, COV = 100%). (From Zhu, H. et al., Computers
andGeotechnics, 48, 249259, 2013. With permission.)

8.4.4.4Groundwater conditions corresponding to the


maximum and minimum factors of safety
Certain hydraulic scenarios under steady-state flow conditions induce the maximum and min-
imum factors of safety. Referring to Figure8.14, the groundwater table which corresponds
to the maximum factor of safety is denoted as curve B, while curve A corresponds to the
minimum factor of safety (the most unfavorable groundwater table). The spatial distributions
of saturated permeability of these two extreme scenarios are displayed in Figure8.18. The
range of ks is from 1.01 106 m/s to 3.53 106 m/s in Figure8.18(a) and from 0.61 104
m/s to 1.85 104 m/s in Figure8.18(b). There is a difference of two orders of magnitude in
ks between these two scenarios although both follow the same prescribed distribution and
correlation structure of ks. Indeed spatially variable permeability function can induce very
different pore-water pressures and groundwater tables, which in turn leads to differences of
critical slip surface.

REFERENCES

Alonzo, E. E., and Krizek, R. J. (1975). Stochastic formulation of soil properties. Proceedings of the
Second Conference on Application of Probability and Statistics to Soil and Structural Engineering,
Vol. 2. Aachen, Germany, pp. 9932.
Asaoka, A., and A-Grivas, D. (1982). Spatial variability of the undrained strength of clays. Journal of
the Geotechnical Engineering Division, ASCE, 108, 743756.
Baecher, G. B. (1982). Simplified geotechnical data analysis. In: Thoft-Christensen, P. (Ed.). Reliability
Theory and Its Applications in Structural and Soil Engineering, Reidal Publishing, Dordrecht, the
Netherlands.
324 Rainfall-Induced Soil Slope Failure

Baecher, G. B., and Christian, J. T. (2003). Reliability and Statistics in Geotechnical Engineering. Wiley,
New York.
Bellin, A., and Rubin, Y. (1996). HYDRO GEN: A spatially distributed random field generator for cor-
related properties. Stochastic Hydrology and Hydraulics, 10(4), 253278.
Benson, C. (1991). Predicting excursions beyond regulatory thresholds of hydraulic conductivity using
quality control measurements. Proceedings of the First Canadian Conference on Environmental
Geotechnics, Montreal, pp. 1417.
Betz, W., Papaioannou, I., and Straub, D. (2014). Numerical methods for the discretization of random
fields by means of the KarhunenLoeve expansion. Computer Methods in Applied Mechanics and
Engineering, 271, 109129.
Bjerg, P.L., Hinsby, K., Christensen, T.H. and Gravesen, P. (1992). Spatial variability of hydraulic con-
ductivity of an unconfined sandy aquifer determined by a mini slug test. Journal of Hydrology,
136(14), 107122.
Bras, R. L., and Rodriyguez-Iturbe, I. (1985). Random Functions and Hydrology. Addison-Wesley
Publishing Company, Reading, MA.
Downloaded by [Monash University Library] at 23:45 02 August 2017

Carpenter, L. C. (1980). Computer Rendering of Fractal Curves and Surfaces. ACM SIGGRAPH
Computer Graphics, 14(3), 915.
Chiasson, P., Lafleur, J., Souli, M., and Haw, K. T. (1995). Characterizing spatial variability of clay by
geostatistics. Canadian Geotechnical Journal, 32(1), 110.
Chiles, J. P., and Delfiner, P. (1999). Geostatistics: Modeling Spatial Uncertainty. Wiley, New York.
Cho, S. E. (2007). Effects of spatial variability of soil properties on slope stability. Engineering Geology,
92(34), 97109.
Cho, S. E. (2012). Probabilistic analysis of seepage that considers the spatial variability of permeability
for an embankment on soil foundation. Engineering Geology, 133134, 3039.
Cooley, J. W., and Tukey, J. W. (1965). An algorithm for the machine calculation of complex Fourier
series. Mathematics of Computation, 19(90), 297301.
Cressie, N., and Pavlicov, M. (2002). Calibrated spatial moving average simulations. Statistical
Modelling, 2(4), 267279.
Davis, M. W. (1987). Production of conditional simulations via the LU triangular decomposition of the
covariance matrix. Mathematical Geology, 19(2), 9198.
DeGroot, D.J. (1996). Analyzing spatial variability of insitu soil properties. Uncertainty in the Geologic
Environment, from Theory to Practice, Proceedings of Uncertainty 96, Madison, WI. Geotechnical
Special Publication (GSP) 58, ASCE, Reston, VA, pp. 210238.
DeGroot, D. J., and Baecher, G. B. (1993). Estimating autocovariance of in-situ soil properties. Journal
of Geotechnical Engineering, ASCE, 119(GT1), 147166.
Deutsch, C. V., and Journel, A. G. (1997). GSLIB: Geostatistical Software Library and Users Guide
(Applied Geostatistics). Oxford University Press, Oxford, 384pp.
Dietrich, C. R., and Newsam, G. N. (1996). A fast and exact method for multidimensional Gaussian
stochastic simulations: extension to realizations conditioned on direct and indirect measurements.
Water Resources Research, 32(6), 16431652.
Ditmars, J.D., Baecher, G.B., Edgar, D.E. and Dowding, C.H. (1988). Radioactive Waste Isolation in
Salt: A Method for Evaluating the Effectiveness of Site Characterization Measurements (Final
report ANL/EES-TM-342), Argonne National Laboratory, Lemont, IL.
Duncan, J. M. (2000). Factors of safety and reliability in geotechnical engineering. Journal of
Geotechnical and Geoenvironmental Engineering, 126(4), 307316.
Emery, X. (2008). A turning bands program for conditional co-simulation of cross-correlated Gaussian
random fields. Computers and Geotechnics, 34(12), 18501862.
Fenton, G. A., and Griffiths, D. V. (2008). Risk Assessment in Geotechnical Engineering. Wiley, New York.
Fenton, G. A., and Vanmarcke, E. H. (1990). Simulation of random fields via local average subdivision.
Journal of Engineering Mechanics, 116(8), 17331949.
Fournier, A., Fussell, D., and Carpenter, L. (1982). Computer rendering of stochastic models.
Communications of the ACM, 25(6), 371384.
Fredlund, D. G., Morgenstern, R. A., and Widger, R. A. (1978). The shear strength of unsaturated soils.
Canadian Geotechnical Journal, 15(3), 313321.
Probabilistic assessment of randomly heterogeneous soil slopes 325

Fredlund, D. G., and Xing, A. Q. (1994). Equations for the soil-water characteristic curve. Canadian
Geotechnical Journal, 31(4), 521532.
Ghanem, R. G., and Spanos, P. D. (1991). Stochastic Finite ElementA Spectral Approach. Springer,
New York.
Griffiths, D. V., and Fenton, G. A. (2004). Probabilistic slope stability analysis by finite element. Journal
of Geotechnical and Geoenvironmental Engineering, 130(5), 507518.
Gutjahr, A., Bullard, B., and Hatch, S. (1997). General joint conditional simulations using a fast Fourier
transform method. Mathematical Geology, 29(3), 361389.
Harter, T. (1994). Unconditional and Conditional Simulation of Flow and Transport in Heterogeneous,
Variably Saturated Porous Media. Ph.D. Thesis, University of Arizona, Tucson, Arizona.
Hess, K. M., Wolf, S. H., and Celia, M. A. (1992). Large-scale natural gradient tracer test in sand and
gravel, Cape Cod, Massachusetts: 3. hydraulic conductivity variability and calculated macrodis-
persivities. Water Resources Research, 28(8), 20112027.
Hilldale-Cunningham, C. (1971). A Probabilistic Approach to Estimating Differential Settlement, M.S.
Thesis, Massachusetts Institute of Technology, Cambridge, MA.
Downloaded by [Monash University Library] at 23:45 02 August 2017

Honjo, Y., and Kuroda, K. (1991). A new look at fluctuating geotechnical data for reliability design.
Soils and Foundations, 31(1), 110120.
Huang, S. P., Quek, S. T., and Phoon, K. K. (2001). Convergence study of the truncated KarhunenLoeve
expansion for simulation of stochastic processes. International Journal for Numerical Methods in
Engineering, 52(9), 10291043.
Jaksa, M. B., Yeong, K. S., Wong, K. T., and Lee, S. L. (2004). Horizontal Spatial Variability of Elastic
Modulus in Sand from the Dilatometer. Proceedings of the 9th Australia New Zealand Conference
on Geomechanics, Vol. 1. Auckland, pp. 289294.
JCSS (2002). Probabilistic Model Code. Joint Committee on Structural Safety. Technical University of
Denmark, Lyngby, Denmark.
Jiang, S. H., Li, D. Q., Cao, Z. J., Zhou, C. B., and Phoon, K. K. (2014). Efficient system reliability
analysis of slope stability in spatially variable soils using Monte Carlo simulation. Journal of
Geotechnical and Geoenvironmental Engineering, 141(2), 04014096.
Jiang, S. H., Li, D. Q., Zhang, L. M., and Zhou, C. B. (2014). Slope reliability analysis considering spa-
tially variable shear strength parameters using a non-intrusive stochastic finite element method.
Engineering Geology, 168, 120128.
Keaveny, J. M., Nadim, F., and Lacasse, S. (1989). Autocorrelation functions for offshore geotechnical
data. Proceedings of the Fifth International Conference on Structural Safety and Reliability, Vol. 1.
Vancouver, BC, pp. 263270.
Kottegoda, N. T., and Kassim, A. H. M. (1991). The turning bands method with the fast-Fourier
transform as an aid to the determination of storm movement. Journal of Hydrology, 127(14),
5569.
Kozintsev, B. (1999). Computations with Gaussian Random Fields. Ph.D Thesis, University of
Maryland.
Kulatilake, P. H. S. W. and Ghosh, A. (1988). An investigation into accuracy of spatial variation estima-
tion using static cone penetrometer data. Proceedings of the First International Conference on
Penetration Testing, Orlando, FL, pp. 815821.
Kulhawy, F. H. (1992). On the evaluation of static soil properties. Stability and Performance of Slopes
and Embankment II, GSP 31, 95115. ASCE, Reston, VA.
Lacasse, S., and de Lamballerie, J. Y. N. (1995). Statistical treatment of CPT data. Proceedings of
International Symposium on Cone Penetration Testing, Linkoping, Sweden, pp. 369380.
Le, T. M. H. (2014). Reliability of heterogeneous slopes with cross-correlated shear strength param-
eters. Georisk: Assessment and Management of Risk for Engineered Systems and Geohazards,
8(4), 250257.
Le, T. M. H., Gallipoli, D., Sanchez, M., and Wheeler, S. J. (2012). Stochastic analysis of unsatu-
rated seepage through randomly heterogeneous earth embankments. International Journal for
Numerical and Analytical Methods in Geomechanics, 36(8), 10561076.
Le, T. M. H., Gallipoli, D., Sanchez, M., and Wheeler, S. (2013). Rainfall-induced differential settle-
ments of foundations on heterogeneous unsaturated soils. Gotechnique, 63(15), 13461355.
326 Rainfall-Induced Soil Slope Failure

Leong, E. C., and Rahardjo, H. (1997). Permeability functions for unsaturated soils. Journal of
Geotechnical and Geoenvironmental Engineering, ASCE, 123(12), 11181126.
Li, D. Q., Jiang, S. H., Cao, Z. J., Zhou, W., Zhou, C. B., and Zhang, L. M. (2015). A multiple response-
surface method for slope reliability analysis considering spatial variability of soil properties.
Engineering Geology, 187, 6072.
Li, D. Q., Qi, X. H., Phoon, K. K., Zhang, L. M., and Zhou, C. B. (2014). Effect of spatially variable
shear strength parameters with linearly increasing mean trend on reliability of infinite slopes.
Structural Safety, 49, 4555.
Li, W., Lu, Z., and Zhang, D. (2009). Stochastic analysis of unsaturated flow with probabilistic colloca-
tion method. Water Resources Research, 45, W08425, DOI:10.1029/2008WR007530.
Lu, Z., and Zhang, D. (2007). Stochastic simulations for flow in nonstationary randomly heterogeneous
porous media using a KL-based moment-equation approach. Multiscale Modeling & Simulation,
6(1), 228245.
Mantolou, A., and Wilson, J. L. (1982). The turning bands method for simulation of random fields
using line generation by a spectral method. Water Resources Research, 18(5), 13791394.
Downloaded by [Monash University Library] at 23:45 02 August 2017

Matrn, B. (1986). Spatial Variation, 2nd ed. Springer-Verlag, Berlin.


Matheron, G. (1973). The intrinsic random functions and their applications. Advances in Applied
Probability, 5(3), 439468.
Meftah, F., Dal Pont, S., and Schrefler, B.A. (2012). A three-dimensional staggered finite element
approach for random parametric modeling of thermo-hygral coupled phenomena in porous media.
International Journal for Numerical and Analytical Methods in Geomechanics, 36(5), 574596.
Mejia, J., and Rodriguez-Iturbe, I. (1974). On the synthesis of random field sampling from the spec-
trum: An application to the generation of hydrologic spatial processes. Water Resources Research,
10(4), 705711.
Mousavi Nezhad, M., Javadi, A. A., and Abbasi, F. (2011). Stochastic finite element modelling of water
flow in variably saturated heterogeneous soils. International Journal for Numerical and Analytical
Methods in Geomechanics, 35(12), 13891408.
Mulla, D. J. (1988). Estimating spatial patterns in water content, matric suction and hydraulic
Conductivity. Soil Science Society, 52(6), 15471553.
Oliver, D. S. (1995). Moving averages for Gaussian simulation in two and three dimensions.
Mathematical Geology, 27(8), 939960.
PDE Solutions Inc. (2015). FlexPDE 6 User Manual Version 6.37. PDE Solutions Inc. Washington, USA.
Phoon, K. K., Huang, S. P., and Quek, S. T. (2002a). Implementation of KarhunenLoeve expansion
for simulation using a wavelet-Galerkin scheme. Probabilistic Engineering Mechanics, 17(3),
293303.
Phoon, K. K., Huang, S. P., and Quek, S. T. (2002b). Simulation of second-order processes using
KarhunenLoeve expansion. Computers & Structures, 80(12), 10491060.
Phoon, K. K., and Kulhawy, F. H. (1996). On quantifying inherent soil variability. Uncertainty in the
Geologic Environment, from Theory to Practice, Proceedings of Uncertainty 96, Madison, WI.
Geotechnical Special Publication (GSP) 58, ASCE, Reston, VA, pp. 326340.
Phoon, K. K., and Kulhawy, F. H. (1999). Characterization of geotechnical variability. Canadian
Geotechnical Journal, 36(4), 612624.
Popescu, R., Prevost, J. H., and Vanmarcke, E. H. (1995). Numerical simulations of soil liquefaction using
stochastic input parameters. Proceedings of the 3rd International Conference on Recent Advances
in Geotechnical Earthquake Engineering and Soil Dynamics, St. Louis, MO, pp.275280.
Rehfeldt, K. R., Boggs. J. M., and Gelhar, L. W. (1992). Field study of dispersion in a heterogeneous
aquifer, 3-d geostatistical analysis of hydraulic conductivity. Water Resources Research, 28(12),
33093324.
Ronold, M. (1990). Random field modeling of foundation failure modes. Journal of Geotechnical
Engineering, 166(4), 554570.
Rosenbaum, M. S. (1987). The use of stochastic models in the assessment of a geological database.
Quarterly Journal of Engineering Geology & Hydrogeology, 20(1), 3140.
Santoso, A. M., Phoon, K. K., and Quek, S. T. (2011). Effects of soil spatial variability on rainfall-
induced landslides. Computers and Structures, 89(1112), 893900.
Probabilistic assessment of randomly heterogeneous soil slopes 327

SoilVision System Ltd (2010). SVSlope Users Manual. SoilVision System Ltd, Saskatoon, Canada.
Souli, M., Montes, P., and Silvestri, V. (1990). Modeling spatial variability of soil parameters. Canadian
Geotechnical Journal, 27(5), 617630.
Srivastava, A., Sivakumar Babu, G. L., and Haldar, S. (2010). Influence of spatial variability of perme-
ability property on steady state seepage flow and slope stability analysis. Engineering Geology,
110(34), 93101.
Stefanou, G. (2009). The stochastic finite element method: past, present and future. Computer Methods
in Applied Mechanics and Engineering, 198(9), 10311051.
Stefanou G., and Papadrakakis M. (2007). Assessment of spectral representation and KarhunenLoeve
expansion methods for the simulation of Gaussian stochastic fields. Computer Methods in Applied
Mechanics and Engineering, 196(21), 24652477.
Strebelle, S. (2002). Conditional simulation of complex geological structures using multiple-point sta-
tistics. Mathematical Geology, 34(1), 121.
Tami, D., Rahardjo, H., and Leong, E. C. (2004). Effects of hysteresis on steady-state infiltration in
unsaturated slopes. Journal of Geotechnical and Geoenvironmental Engineering, 130(9), 956967.
Downloaded by [Monash University Library] at 23:45 02 August 2017

Tang, W. H. (1979). Probabilistic evaluation of penetration resistances. Journal of the Geotechnical


Engineering Division, ASCE, 105(GT10), 11731191.
Tang, W. H. (1984). Principles of probabilistic characterization of soil properties. Proceedings of
Symposium on Probabilistic Characterization of Soil Properties, ASCE National Convention,
Atlanta, pp. 7489.
Tartakovsky, D. M., Lu, Z., Guadagnini, A., and Tartakovsky, A. M. (2003). Unsaturated flow hetero-
geneous soils with spatially distributed uncertain hydraulic parameters. Journal of Hydrology,
275(34), 182193.
Tompson, A. F. B., Ababou, R., and Gelhar, L. W. (1989). Implementation of the three-dimensional turn-
ing bands random field generator. Water Resources Research, 25(10), 22272243.
nl, K., Nielsen, D. R., Biggar, J. W., and Morkoc, F. (1990). Statistical parameters characterizing the
spatial variability of selected soil hydraulic properties. Soil Science Society of America Journal,
54(6), 15371547.
Vanmarcke, E. (1983). Random Fields: Analysis and Synthesis. MIT Press, Cambridge, MA.
Vanmarcke, E. H. (1977). Probabilistic modeling of soil profiles. Journal of the Geotechnical Engineering
Division, ASCE, 103(GT11), 12271246.
Vrouwenvelder, T., and Calle, E. (2003). Measuring spatial correlation of soil properties. Heron, 48(4),
297311.
Whitman, R. V. (2000). Organizing and evaluating uncertainty in geotechnical engineering. Journal of
Geotechnical and Geoenvironmental Engineering, 126(7), 583593.
Wu, T. H. (1974). Uncertainty, safety, and decision in soil engineering. Journal of Geotechnical
Engineering, 100(3), 329348.
Yaglom, A. M. (1962). An Introduction to the Theory of Stationary Random Functions. Dover
Publications Inc., Mineola, NY.
Yaglom, A. M. (1987). Correlation Theory of Stationary and Related Random Functions. Vol. I: Basic
Results. Springer, Berlin.
Yang, J., Zhang, D., and Lu, Z. (2004). Stochastic analysis of saturatedunsaturated flow in hetero-
geneous media by combining Karhunen-Loeve expansion and perturbation method. Journal of
Hydrology, 294(1), 1838.
Zhang, D. (1999). Nonstationary stochastic analysis of transient unsaturated flow in randomly hetero-
geneous media. Water Resources Research, 35(4), 11271141.
Zhang, L. L., Fredlund, D. G., Zhang, L. M., and Tang, W. H. (2004). Numerical study of soil conditions
under which matric suction can be maintained. Canadian Geotechnical Journal, 41(4), 569582.
Zhang, L. M., Zheng, Y. R., and Wang, J. L. (2003). Errors in calculating hydrodynamic pressures for
stability analysis of soil slopes subject to rainfall. Proceedings of ICASP9, San Francisco, CA, pp.
14311438.
Zhu, H., Zhang, L. M., Zhang, L. L., and Zhou, C. B. (2013). Two-dimensional probabilistic infiltra-
tion analysis with a spatially varying permeability function. Computers and Geotechnics, 48,
249259.

S-ar putea să vă placă și