Sie sind auf Seite 1von 12

Computers and Geotechnics 110 (2019) 296–307

Contents lists available at ScienceDirect

Computers and Geotechnics


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

Research Paper

Three-dimensional seismic stability of unsaturated soil slopes using a semi- T


analytical method
Long Wanga,c, De'an Suna, , Bo Chenb, Jie Lic

a
Department of Civil Engineering, Shanghai University, Shanghai 200444, China
b
College of Civil Engineering and Architecture, Quzhou University, Zhejiang 324000, China
c
School of Engineering, RMIT University, Melbourne 3001, Australia

ARTICLE INFO ABSTRACT

Keywords: Seismic stability analyses of soil slopes are routinely performed with dry and/or saturated assumptions under
Seismic stability two-dimensional (2D) conditions. However, in practice, soils are often unsaturated and slopes usually fail in 3D
Limit analysis fashion. Three-dimensional (3D) effect and suction-induced effect are essential to be considered to produce more
Unsaturated soil realistic solutions in the slope stability analyses. In this regard, an analytical framework is developed in this
Gentle soil slope
paper for assessing the 3D seismic stability of unsaturated soil slopes. A semi-analytical method including the
Semi-analytical method
simplified method (SM) and the layer-wise summation method (LSM) is developed to deal with the highly
variable nature of suction. The validity of the analytical framework in assessing the soil slope stability is checked
with published solutions. Examples are given to illustrate the 3D effect and the effect of suction on the slope
seismic stability. The described framework allows the user to quantify the effect of changes in moisture content
on effective stress distribution and effective unit weight distribution in the slope and thus, changes in the slope
stability.

1. Introduction LEMs (e.g., [14,47]) are still the classical approaches, and can be used
to the safety estimation of 3D slopes by extending the 2D cases. How-
Earthquake-related damages to the soil slopes or dams have been ever, the method is statically indeterminate as arbitrary assumptions
widely reported and investigated (e.g., [25,8]). Seismic excitation is are introduced to determine the location of the thrust line or surface to
one of the possible factors leading to failures of slopes or dams. Stability satisfy the force and moment equilibriums. Furthermore, the results
assessments of slopes subjected to seismic shaking are routinely per- given by the LEMs are neither a strict lower- nor upper-bound solutions
formed under two-dimensional (2D) conditions using a pseudo-static [19]. The numerical approaches, such as the finite element method
approach (e.g., [2,46]). In the pseudo-static approach, the seismic ex- (FEM) (e.g., [7,17]), etc., have gained significant attention in geo-
citation caused by seismic acceleration is treated as a steady uniformly technical engineering. FEM is more applicable for specific, well-defined
distributed inertial force (horizontal and/or vertical). Though this ap- slopes with accurate geotechnical parameters, but would become more
proach does not reflect the characteristics of seismic excitation (e.g., the elaborate and time-consuming to produce solutions for a wide range of
duration, periodicity and amplification, etc.) on the slope stability, it parameters. Compared with the LEM, the classical plasticity theorem-
allows for implementing a quantitative assessment of the slope stability based LAM proposed originally by Drucker and Prager [6] is a much
and is widely used in the preliminary stage for designing new slopes or more rigorous method for assessing the stability of geostructures (e.g.,
the safety assessment of existing slopes. For slopes failing with three- [3,24,35]). Meanwhile, with the advent and development of 3D ad-
dimensional (3D) fashion, the 2D approaches may be overly con- missible failure mechanisms in limit analysis (e.g., [5,27,10]), it is
servative (e.g., [7,4]). To produce more realistic solutions, it is of possible to evaluate the 3D seismic stability of slopes using the LAM.
practical significance to perform 3D stability analyses of slopes. In this Michalowski and Martel [28] adopted the pseudo-static approach to
regard, three main approaches, i.e., the traditional limit equilibrium assess the 3D seismic stability of steep soil slopes assuming a toe-failure
method (LEM), the robust numerical approaches and the limit analysis mechanism. Nadukuru and Michalowski [29] implemented a 3D dis-
method (LAM) can be used to assess the 3D seismic stability of slopes. placement analysis of slopes subjected to seismic loads by adopting a
The above approaches are used for their respective advantages. The below-toe failure mechanism. Zhang et al. [45] investigated the effect


Corresponding author at: 99 Shangda Road, Shanghai 200444, China.
E-mail address: sundean@shu.edu.cn (D. Sun).

https://doi.org/10.1016/j.compgeo.2019.02.008
Received 12 October 2018; Received in revised form 16 January 2019; Accepted 8 February 2019
0266-352X/ © 2019 Elsevier Ltd. All rights reserved.
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

of vertical seismic acceleration on the 3D slope stability and presented


the stability charts. They found that the slope stability can be over-
estimated by more than 10% when the vertical seismic acceleration is
ignored in assessing the 3D seismic stability of slopes. Xu and Yang [43]
conducted a 3D stability analysis of nonhomogeneous and anisotropic
slope subjected to seismic force. Their results showed that, the seismic
force will not only influence the slope stability but also change the rules
of soil nonhomogeneity and anisotropy on slope stability.
The aforementioned works most adopted the fully dry and/or sa-
turated assumptions to evaluate the slope stability. However, most
slopes are in an unsaturated state and the fully dry and/or saturated
assumptions cannot account for the stabilizing effect of moisture or
suction. The frequent occurrences of precipitation-induced failures to
unsaturated soil slopes (e.g., [15,33]) highlight the significance for
explicit consideration of suction in the slope stability analyses. Li and
Yang [20] investigated the 3D stability of unsaturated soil slopes sub-
jected to unsaturated steady flow using the LAM. However, in their
Fig. 2. Shear strength surface for unsaturated soils (redrawn after [23]).
work, they neglected the energy dissipated within the soil volume due
to the capillary cohesion DVs and the effect of moisture content on soil’s
unit weight in calculating the external work W . This may be due in a 2. Shear strength of unsaturated soils
large part to the fact that, the highly variable nature of suction (e.g.,
[23,40]) brings difficulty in deriving an upper bound solution based on It has been shown that the capillary cohesion due to suction may
the upper bound theorem in limit analysis. contribute substantially to the stability of natural and artificial soil
The 3D static stability of unsaturated soil slopes has been studied in slopes (e.g., [16,18]). Based on a semi-quantitative validation of the
our previous publications [41,42], where the validities of the layer-wise suction stress data from 14 different soils, Lu et al. [22] presented an
summation method (LSM) were demonstrated by comparing with equation of the extended Mohr-Coulomb failure criterion, that is, the
published solutions of Gao et al. [10]. To explore the effect of seismic effect of suction stress is treated as a cohesion value. Graphical re-
forces on the slope stability, this paper describes an analytical frame- presentation of the extended Mohr-Coulomb criterion is plotted in a 3D
work for assessing the 3D seismic stability of unsaturated soil slopes. stress space as shown in Fig. 2. It can be seen form Fig. 2 that the failure
For this, a 3D rotational failure mechanism first proposed by Micha- surface is planar in the stress space. This feature indicates that, the
lowski and Drescher [27] (Fig. 1) for no suction case is extended to an intercept c′ in classical Mohr-Coulomb criterion can be equivalently
unsaturated condition. The simplified method (SM) and the LSM are replaced by a new intercept c′ + c′′ to reflect the effect of suction on the
further developed to deal with the highly variable nature of suction in shear strength. Hence, the classical theories of soil mechanics for sa-
gentle soil slopes. A brief description of the upper bound theorem of turated conditions can be readily extended to unsaturated conditions
limit analysis is presented in the next section, followed by the de- (e.g., [12,40,20]). Based on the suction stress characteristic curve, the
scription and verification of the semi-analytical method in assessing the shear strength of unsaturated soils can be expressed using effective
3D seismic stability of unsaturated soil slopes. The paper is concluded stress parameters (c′ and ϕ′) as [22]
with examples and conclusions.
f =c s tan +( ua ) tan
=c +c +( ua ) tan (1)

where f depicts the shear strength of unsaturated soils; c′ and ϕ′ are the
effective cohesion and internal friction angle, respectively; s represents
the suction stress; ua denotes the net stress on the failure plane; and
ua is the pore-air pressure and is assumed to be zero in this paper.
The second term on the right-hand side in Eq. (1), c = s tan ,
describes the shearing resistance arising from capillary effects and is
usually defined as capillary cohesion. The suction stress primarily arises
from the combination of the interparticle physicochemical stresses and
the capillary stresses caused by surface tension and negative pore-water
pressure. The equation for s can be expressed as a function of matric
suction (ua u w ) with two fitting parameters α and n and is given as
follows [22]:

(u a uw )
s = ua uw 0
{1 + [ (ua u w )]n }(n 1) n
(2)

where uw is the pore-water pressure; α approaches the inverse of the air-


entry pressure (typically, 0 ≤ α ≤ 0.5 kPa−1); and n describes the pore-
size distribution (typically, 1.1 ≤ n ≤ 8.5) [38,23]. Eq. (2) can be ap-
plied to both wetting and drying processes by adjusting the parameters
α and n. Likos et al. [21] investigated the effect of hysteresis on the
hydrologic and mechanical responses of unsaturated soils, and found
that the parameter n is almost constant for both wetting and drying
paths for different soils and the average ratio αw/αd is 2.24 ± 1.25 (the
Fig. 1. Schematic of 3D rotational failure mechanism based on the concept of superscripts w and d indicate wetting and drying paths, respectively).
Michalowski and Drescher [27]. The hydrologic and mechanical parameters for different soils under

297
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

Table 1
Parameters for different soils under drying and wetting conditions (after [21]).
Soil USCS c' (kPa) ϕ′ (°) n α (kPa−1) Sr ksat (10–7 m/s) Gs

D W D W D W D W

A SC-SM 20.3 32 1.75 1.75 0.05 0.05 0.342 0.433 1.20 0.10 2.70
B SM 8.8 24 1.44 1.52 0.12 0.13 0.227 0.303 1.10 0.039 2.70
C ML 12.5 35 1.50 1.46 0.12 0.13 0.319 0.375 8.00 0.059 2.70
D SM 6.6 37 2.00 2.20 0.08 0.17 0.380 0.413 1.00 0.146 2.70

Note: D = drying; W = wetting; For lack of data in situ, the specific gravity (Gs) of each soil is assumed to be 2.70 in this analysis.

10 10
Distance from water table, z (m)

Distance from water table, z (m)


8 8

6 q/ks 6 q/ks
0 0
4 - 0.2 4 - 0.2
- 0.4 - 0.4
- 0.6 - 0.6
2 - 0.8 2 - 0.8
- 1.0 (NS) - 1.0 (NS)

0 Water table 0 Water table 3


c'' = 0 sat
= 20 kN/m
0 5 10 15 20 25 18.0 18.5 19.0 19.5 20.0
3
Capillary cohesion, c'' (kPa) Effective unit weight, (kN/m )
(a) (b)
Fig. 3. Characteristics of an unsaturated soil layer with a depth of 10 m for Soil A: (a) capillary cohesion profile; and (b) effective unit weight profile (NS represents
no suction case).

drying and wetting conditions are listed in Table 1 [21]. Note that, for 1 ln[(1 + q ks ) e wz q ks ]
s =
unsaturated soils, the suction (or water content) obtained along any {1 + { ln[(1 + q k s ) e wz q k s ]}n}(n 1) n
(4)
wetting path is generally smaller than that obtained at the same water
In application of limit analysis to slopes without surface loads, the
content (or suction) along a drying path. This means that the effect of
self-weight of soil is usually identified as the external force. When the
suction on the slope stability is less pronounced when the soil para-
effect of suction is included in the slope safety assessment, the variation
meters along wetting paths are utilized in the slope stability analyses.
of moisture content would undoubtedly alter the unit weight of soils,
Hence, to yield a conservative estimation of the slope stability, only the
and thus affects the slope stability. In routine slope stability analyses,
wetting path is considered in present analysis.
the variation of moisture content is usually neglected (e.g., [40,30,20]),
Combining transient suction analysis with the LAM may bring dif-
which may yield unrealistic results. According to the proportional re-
ficulties in performing a 3D stability analysis of unsaturated soil slopes.
lationship between the gases, liquid and solid phases of unsaturated
As a preliminary study, this paper focuses on the stability of un-
soils, the effective unit weight of unsaturated soils can be derived as
saturated soil slopes under steady flow conditions. Combining Darcy’s
law and Gardner’s [11] model and considering the flow boundary ( sat w ) Gs + [Se (1 Se ) Sr ](Gs w sat )
=
conditions (zero suction at water table elevation), the matric suction Gs 1 (5)
versus depth relation under vertical steady flow condition can be ob-
tained analytically as follows [44,23]: where γsat represents the unit weight of saturated soil; Gs denotes spe-
cific gravity of soil; Se and Sr depicts the effective and residual degrees
1 q q of saturation, respectively. For most inorganic silts and clays, the value
(u a u w) = ln 1+ e wz
of Gs varies over a narrow range of 2.60 ∼ 2.80 [32]. For lack of data in
ks ks (3)
situ, the values of γsat and Gs, which can be easily ascertained in a la-
where q denotes the steady flow rate (negative value for infiltration, boratory or in situ, are assumed to be equal to 20 kN/m3 and 2.7 in this
positive value for evaporation and zero for hydrostatic condition); ks study.
describes the saturated hydraulic conductivity; γw is the unit weight of By using the soil water characteristic curve by Van Genuchten [38],
water; and z represents a distance from a generic point within the slope the relationship between the effective degree of saturation and matric
to the water table level. Eq. (3) implies that the matric suction in a soil suction can be written as
slope varies only in the vertical direction and is constant in the same 1
1 1 n

horizontal plane. Though the possible effect of transient flow near the Se =
1 + [ (u a uw )]n (6)
slope surface and the effect of slope shape on the suction distribution
cannot be considered, the vertical steady flow assumption allows for Fig. 3 shows the characteristics of Soil A under different infiltration
making a tractable analysis of the slope stability and is widely adopted conditions. The normalized flow rate q/ks is used and an unsaturated
in conventional slope stability analyses (e.g., [16,40,20]). soil layer with a depth of 10 m is investigated. Fig. 3(a) depicts the
Substituting Eq. (3) into Eq. (2), the suction stress profile under capillary cohesion profile and Fig. 3(b) describes the effective unit
vertical steady flow condition can be expressed as weight profile, respectively. As seen from Fig. 3, both the capillary

298
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

cohesion profile and the effective unit weight profile exhibit nonlinear
shapes. Moreover, as the infiltration rate increases, the capillary co-
hesion profile decreases continually and approaches no suction case
(i.e., NS corresponds to q/ks = −1.0 in Eq. (4)) and the effective unit
weight profile increases continually and approaches that of saturated
condition (i.e., γsat = 20 kN/m3). More detailed description about suc-
tion can be found in the references (e.g., [23,22]).

3. Kinematic approach of limit analysis theorem

The kinematic approach of limit analysis theorem is a well-estab-


lished method in dealing with the stability problems in geotechnical
engineering [35]. For a kinematically admissible failure mechanism, a
balance equation of work rate can be given and an upper bound solu-
tion can be derived from the energy balance equation. In application of
the kinematic approach of limit analysis theorem to unsaturated soil Fig. 5. Transition to 2D failure mechanism (b → ∞) with a plane insert of width
slopes, it is assumed that the unsaturated soil deforms plastically fol- b based on the concept of Michalowski and Drescher [27].
lowing the normality rule associated with the extended Mohr-Coulomb
yield criterion using the effective strength parameters (c′ and ϕ′). failure surface, a hypothetical angle β′ is introduced. When the angle β′
For slopes with smaller inclination angles, which are called gentle is increased to an angle β, a 3D toe-failure mechanism is obtained.
slopes, the slip surfaces usually pass below the slope toe and intersect Though the variation of the water table highly depends on the site-
the slope crest. To investigate the 3D stability of gentle unsaturated soil specific climate and soil properties, it can nearly be neglected when the
slopes using the kinematic approach of limit analysis theorem, a kine- slope is momently subjected to an earthquake excitation. In this study,
matically admissible 3D rotational failure mechanism (e.g., [29]) is the water table is assumed be horizontal and is located below the slope
extended to an unsaturated condition. Fig. 4 shows the 3D ‘horn-shape’ toe with a depth of z0. As suggested by Michalowski and Drescher [27],
rotational failure mechanism for a gentle unsaturated soil slope in the 3D failure mechanisms are separated laterally to incorporate a
frictional/cohesive soils (ϕ' > 0 and c' > 0). The ‘horn-shape’ me- plane insert of width b as shown in Fig. 5. The 2D failure mechanism [3]
chanism can be constructed by rotating a circle of increasing radius R is obtained as b approaches infinity. This treatment allows for im-
about an axis rm passing through point O outside the circle. An alter- plementing a tractable comparison of the 3D effect on the slope stabi-
native mechanism can also be generated when the circle is rotated lity. Detailed description about the 3D rotational failure mechanism can
about an axis passing through the circle. Graphic illustration about the be found in the references (e.g., [27,10,29]).
alternative mechanism and the 3D rotational failure mechanisms in To account for the effect of seismic load on the slope stability, an
purely cohesive soils (ϕ′ = 0° and c′ > 0) can be found in Michalowski additional rate of work done by the pseudo-static seismic forces must be
and Drescher [27] and is not repeated here. The surface of the ‘horn- included in the energy balance equation. For a given kinematically
shape’ is smooth and has a symmetry plane. Only a portion of this admissible failure mechanism, an upper bound solution can be derived
surface intersects the slope (see the yellow area of the slope in Fig. 4). from an inequality equation of work rate, which is given as
The bottom trace of the mechanism on the symmetry plane passes
below the slope toe and intersects the slope crest. The sliding soil mass
(soil block ABCD) can be modeled by a single rigid rotational block with k k
ij ij dV Wi( ) vik dV + Wi(s) vik dV
rotation axis O and angular velocity ω. To determine the geometry of V V V (7)

Fig. 4. 3D ‘horn-shape’ rotational failure mechanism for a gentle unsaturated soil slope based on the concept of Nadukuru and Michalowski [29].

299
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

Fig. 6. Calculation of the external work rate done by unsaturated soil weight: (a) for 2D plane insert; and (b) for 3D failure mechanism.

where k
ij and k
ij represent the stress and the strain rate tensor in a slope stability [45]. When kv ≤ 0.5 kh, the effect of vertical seismic
kinematically admissible mechanism, respectively; Wi( ) and Wi(s) de- acceleration on the 3D slope stability can nearly be ignored (e.g.,
scribe the work rates done by the gravity force and seismic force, re- [13,45]). Thus, only the effect of horizontal seismic acceleration on the
spectively; vik depicts the velocity vector consistent with the kinemati- 3D slope stability is considered in this paper.
cally admissible strain ijk ; and V is the volume of the rotating block. Andrianopoulos et al. [1] and Papadimitriou et al. [31] presented a
methodology for estimating the horizontal seismic coefficient kh of
earth dams and tall embankments. They suggested using an ‘‘effective’’
4. A semi-analytical method
value of the horizontal seismic coefficient, khE (approximately, khE/
khmax = 0.67 based on their literature), instead of the peak value of the
In this study, the external work rate is caused by the soil’s self-
seismic coefficient khmax in the design process. This is because that
weight and seismic load, while the internal energy is dissipated due to
using the value of khmax yields an overly conservative estimation of the
the soil cohesion and capillary cohesion. The energy balance equation
stability of geostructures. Hence, in this paper, only the ‘‘effective’’
can be expressed as
horizontal seismic coefficient khE is used in assessing the slope stability.
W + Ws = Dc' + Dc'' (8) Note that, for gentle soil slope subjected to stronger seismic ex-
citation, the critical failure surfaces tend to be very large and deep, and
where W and Ws represent the external work rates done by the un-
khE > 0.2 from Kramer [26]. To produce a rational result, as proposed
saturated soil weight and the seismic load, respectively; and Dc' and Dc''
by Gao et al. [9], the depth of the below-toe failure mechanism is
are the internal energy dissipation rates caused by the soil cohesion and
limited to a realistic value D below the slope toe elevation as shown in
the capillary cohesion, respectively. The expression of Dc' is available in
Fig. 4. In this paper, the depth factor D/H = 1.0 is used for the slopes
Nadukuru and Michalowski [29]. In the following, only detailed cal-
subjected to stronger seismic excitation.
culations of W , Ws and Dc'' are involved. Note that, in this paper the
effects of seismic excitation and infiltration on the slope stability are
4.2. A SM to calculate the external power rate
uncoupled. This will greatly facilitate the development of the analytical
framework.
Because of the highly variable nature of the effective unit weight
profile (Fig. 3b), it is hard to calculate the external work rate of the
4.1. Pseudo-static analysis unsaturated soil weight using a direct integrate method. Based on the
volume integral of the rotating body, a SM is developed to calculate the
The pseudo-static approach, though approximate, has the ad- work rate of the unsaturated soil weight. The volume element can be
vantages of accumulating experience and user friendliness and is widely written as
used in evaluating the seismic stability of geotechnical structures since
dV = A ( ) lc d (9)
it was first employed in 1950s [37]. Fig. 4 shows some significant
parameters of this approach, such as the peak values of the seismic where A(θ) is the area of soil represented by the line segment
acceleration at the outcropping bedrock, PGArock, at the ground surface, ¯ (m¯n , m ¯n ) for the plane insert (Fig. 6a) or the area of soil inter-
mn
PGA and at the slope crest, PGAcrest. In this approach, the seismic ex- sected by the curvilinear cone surface (the shaded region in Fig. 6b). lc
citation is treated as a uniformly distributed inertial force, Fh and Fv (h is a distance from the center point of area A(θ) to the rotation axis O.
and v denote the horizontal and vertical directions), which are applied The rotation velocity v regarding to the center point of area A(θ) is
at the sliding soil mass’s center of gravity as shown in Fig. 4. Hence, the perpendicular to the distance lc and can be given as
most important point of this approach is to determine the values of Fh
v = lc (10)
and Fv, which can be calculated by multiplying the weight of soil block,
Wγ', with a dimensionless seismic coefficients kh and kv, respectively. where ω is the angular velocity.
Vahedifard et al. [39] investigated the relationship between the To make the calculation manageable, the local nonlinear feature of
seismic coefficients kh and kv on the 2D stability of reinforced earth the effective unit profile in the volume element dV is neglected. The
structures. They found that when the direction of vertical seismic ac- weight of the volume element is calculated approximately by multi-
celeration is consistent with the direction of gravity, the stability of plying the volume and the effective unit weight at the center point (i.e.,
reinforced earth structures is enhanced. However, for a 3D approach, c (c′, c'') in Fig. 6). This assumption, though approximate, allows for
the vertical seismic acceleration with a downward direction reduces the implementing a relatively simple calculation of the external work rate

300
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

done by the unsaturated soil weight. This is because that, in practice, cj = 0.5[ s (z )|z = jh1 + s (z )|z = (j 1) h1 ] tan (j = 1, 2 …m1) (14)
slopes or dams are usually constructed by fine-grained soils and the unit
weight profiles of soils generally present a linear distribution. The and
overall external work rate (including the work rate done by the soil ck = 0.5[ s (z )|z = (z 0 +
N ) + kh2 s (z )|z = (z 0 N ) + (k 1) h2 ] tan (k = 1,
weight and seismic load) for soil block ABCD (including the 2D and 3D
cases) are given as follows: 2 …m2) (15)
where s (z )|z = jh1 and s (z )|z = (j 1) h1 describe the suction stress at the top
W 2D + Ws2D
and bottom surfaces of layer j for soil block ABCG (Fig. 7b), respec-
1
= 4 br03 { 0
B
(z )|z = z1
(cos + k hE sin )
sin3
[sin e( 0 ) tan
+ sin 0]
2 tively; s (z )|z = (z 0 N ) + kh2 and s (z )|z = (z 0 N ) + (k 1) h2 capture the suction
stress at the top and bottom surfaces of layer k for soil block CDG
× [sin e ( 0) tan
sin 0]d (Fig. 7c), respectively. The expressions of s (z )|z = jh1, s (z )|z = (j 1) h1,
s (z )|z = (z 0 N ) + kh2 and s (z )|z = (z 0 N ) + (k 1) h2 can be easily got by sub-
C (cos + k hE sin )
+ (z )|z = z2 3
sin ( + ) sin3 B
stituting the variable z in Eq. (4) with jh1, (j − 1)h1, (z0 - N) + kh2 and
B
( 0 ) tan 2
[sin( + ) sin Be + sin( B + ) sin 0] (z0 - N) + (k − 1)h2, respectively. Note that, when the water table
× [sin( + ) sin Be
( 0 ) tan
sin( B + ) sin 0 ]d intersects the slip surface (i.e., z0 < N), the capillary cohesion below
(cos + k hE sin ) the water table is zero and the effective unit weight is equal to saturated
+ h
(z )|z = z3 [sin e( 0 ) tan
+ sin he
( h 0 ) tan
]2
C sin3 unit weight.
( ) tan ( h 0 ) tan
× [sin e 0
sin he ]d } For slopes in cohesive-frictional soils (ϕ′ > 0° and c′ > 0), the
(11) overall dissipated work rate due to the capillary cohesion of each soil
layer (e.g., layer j) (sum of the dissipation within volume DV and that
W 3D + Ws3D over the velocity discontinuity surfaces Dt) can be derived based on the
divergence theorem as
B
= (z )|z = z 4 (cos
0
Dc''3D - j = c j cot Srem - j
vi ni dSrem - j
2 2 2 2
R a R x
+ khE sin )(rm + x c1 )2 dydxd
0 a =c j cot vi ni dS + v n dS
Sj j 1 i i
+ vi ni dS
C R 2
d 2
R2 x 2
Sj j Sj 1j 1 (16)
+ (z )|z = z5 (cos + khE sin )(rm + x c2 )2 0 d
dydxd
where Srem-j = Sj′j + Sjj-1 + Sj-1j′-1 (Sj′j, Sjj-1 and Sj-1j′-1 are the upper
B

R2 e 2 R2 x 2
+ h
(z )|z = z 6 (cos + khE sin )(rm + x c3 ) 2 dydxd surface, slope surface and bottom surface of soil layer j, respectively, as
C 0 e
shown in Fig. 7b).
(12) The dissipation work rate Dc''3D caused by the capillary cohesion in
whole soil block ABCD can be divided into two parts (i.e., the energy
where khE is the effective horizontal seismic coefficient (typically,
dissipated in soil block ABCG, Dc''ABCG - 3D , and that dissipated in soil
0.1 ≤ khE ≤ 0.3) [26], (z )|z = z1, (z )|z = z2 and (z )|z = z3 capture the ef-
block CDG, Dc''CDG - 3D), which are given, respectively, as
fective unit weights corresponding to the center points of the areas
¯ , m¯n and m ¯n , respectively, as
represented by the line segment mn
shown in Fig. 6(a); (z )|z = z 4 , (z )|z = z5 and (z )|z = z 6 represent the ef-
Dc''3D - ABCG = cot { 0
B
c (z )|z = z7 vi ni dS +
B
C
c (z )|z = z 8 vi ni dS

fective unit weights corresponding to the center points of areas A(θ) as


shown in Fig. 6(b); and xc1, xc2 and xc3 are the distances from the center
G
C
c (z )|z = z 0 vi ni dS }
m1 - 1
points of areas A(θ) to the centerline of the conical volume, respec- (c j c j + 1) cot vi ni dS
Sjj
tively, as shown in Fig. 6(b). Expressions of xc1, xc2 and xc3 are given in j=1

Appendix. The expressions of (z )|z = z1, (z )|z = z2 , (z )|z = z3, (z )|z = z 4 ,


(z )|z = z5 and (z )|z = z 6 can be easily obtained according to Eqs. (3), (5)
= 2 r02 sin2 { 0
0
B cos
s (z )|z = z7 sin3 R2 a2 d

and (6) by substituting the variable z with the variables z1, z2, z3, z4, z5
and z6, which are given in Appendix. The variables a, d, e, R, rm, θB and
+
sin2 ( + C) sin2 h 2( h
sin2 C
e 0 ) tan C
B
cos( + )
s (z )|z = z 8 sin3 ( + ) R2 d2 d }
m1 - 1
θC can be easily derived from the geometrical relationships as shown in 2 cot r j20 (c j c j + 1)sin
2
j
j cos
R2 a2j d
sin3
Fig. 4, which are available in the references (e.g., [29]). j =1
j

C
c (z )|z = z 0 cot vi ni dS
4.3. A LSM to calculate the internal energy dissipation rate
G

(17a)
Because of the highly variable nature of the capillary cohesion and
profile (Fig. 3b), a LSM is developed to calculate the internal energy m2
dissipation rate based on the divergence theorem. To make the calcu- Dc''3D - CDG = c k cot { Sk k
vi ni dS + v n dS
Sk 1k 1 i i }
lation manageable, the soil mass is divided into two parts, i.e., soil k=1
cos
block ABCG (above the slope toe elevation) and soil block CDG (below =2 r02sin2 h e 2( h 0) tan h
s (z )|z = z 0 sin3 R2 e2 d
C
the slope toe elevation) as shown in Fig. 7. Soil blocks ABCG and CDG m2 - 1
k cos
are divided into m1 and m2 discrete layers, respectively. The thickness 2 rk20 (c k c k + 1) cot sin2 k sin3
R2 ek2 d
k
k=1
of the soil layer in each part are h1 = H/m1 and h2 = N/m2. The value C
of m2 can be approximately determined according to the relationship + c (z )|z = z 0 cot vi ni dS (17b)
G
between N and H once the value of m1 is given. The relationship be-
where s (z )|z = z7 , s (z )|z = z 8 and s (z )|z = z 0 describe the suction stress at
tween N and H takes the following form
slope crest AB, slope surface BC and ground surface CD, respectively,
N cos e( 2 + 0) tan
sin which can be easily got by substituting the variable z in Eq. (4) with the
0
= ( 1 variables z7, z8 and z0; and r j0 and rk0 are shown in Fig. 7. The ex-
H e h 0 ) tan
sin h sin 0 (13)
pressions of the variables z7, z8, z0, aj, θj′, θj, ek, θk′ and θk are given in
The capillary cohesion of the soil layer in each part (e.g., layer j and Appendix. Note that, the last terms on the right-hand side in Eqs. (17a)
k) can be approximately written, respectively, as and (17b) are opposite in sign and equal in magnitude.

301
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

Fig. 7. LSM to calculate the energy dissipation rate of a gentle unsaturated soil slope: (a) notation and convention for LSM; (b) 3D diagram of Layer j; and (c) 3D
diagram of Layer k.

For the 2D plane insert, the dissipation work rate Dc''2D can be derived factor, which is extremely time-consuming to solve under 3D condi-
as follows: tions. Conversely, the GIM has an advantage in giving an explicit ex-
h
pression of the safety factor, which may be preferred by geotechnical
Dc''2D = br02 tan s (z )|z = z9 e
2( 0) tan
d engineers. Thus, the GIM is adopted in this paper. For the gentle un-
0 (18)
saturated soil slope with a given height, the safety factor Fs can be
where s (z )|z = z9 describes the suction stress at a generic point on the slip expressed as
surface, which can be easily got by substituting the variable z in Eq. (4)
Dc' + Dc''
with the variable z9. The expression of the variable z9 is given in Ap- Fs =
pendix.
W + Ws (20)
For the special case of purely cohesive soils (ϕ′ = 0° and c′ > 0), the The fundamental inequality of the kinematical approach of limit
expression of Dc''3D is given as analysis theorem implies that the safety factor should not less than 1.0
(i.e., the internal energy dissipated rate Dc' + Dc'' is not less than the
Dc''3D = 0 (19)
external power rate W + Ws ) for a slope to be safe. When Fs < 1.0, the
In geotechnical engineering practice, the strength reduction method soil starts yielding and displacements of the sliding soil mass may take
(SRM) (e.g., [48]) and the gravity increase method (GIM) (e.g., [34,30]) place, which means the slope becomes unstable.
are widely used for slope safety estimation. The above two definitions The basic plasticity theorems of limit analysis can be cast as opti-
of the safety factor are valid although not necessarily the same [36]. mization problems. An optimization procedure is developed with in-
Nevertheless, when applied to the kinematical approach of limit ana- dependent variables θ0, θh, r′0/r0 and β′ to find the global minimum
lysis theorem, the SRM gives an implicit equation about the safety value of the safety factor. According to the 3D rotational failure

302
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

Table 2 5.2. Parameter effects on LSM


Comparison of stability factor γH/c of gentle soil slopes obtained by Gao et al.
[10] and this paper. Fig. 8 investigated the effects of internal friction angle ϕ′ (Fig. 8a)
Case B/H γH/c and inclination angle β (Fig. 8b) on calculated results. An unsaturated
clayey soil slope (n = 2.0, α = 0.005 kPa−1, Sr = 0.16) with para-
β = 15°, ϕ′ = 0° β = 30°, ϕ′ = 0° β = 30°, ϕ′ = 15° meters c′ = 20 kPa, z0 = 0.5H, khE = 0 and q = 0 m/s is investigated.
The calculated safety factors and the corresponding Dm1 are also pre-
Gao Present Gao Present Gao Present
et al. study et al. study et al. study sented in Tables 3 and 4, respectively. As expected, the value of Dm1
[10] [10] [10] generally converges to zero as the layer number increases. For example,
for a slope with β = 30°, H = 10 m, B/H = 2.0 and ϕ′ = 10° (Fig. 8a),
1 1.0 17.312 17.114 11.885 11.875 38.011 39.525
the deviation Dm1 is 0.25% for m1 = 10 and is reduced to 0.09% for
2 2.0 11.853 11.821 8.615 8.369 28.098 28.095
3 3.0 9.108 9.430 7.417 7.472 25.544 25.542 m1 = 30.
4 5.0 7.796 7.815 6.713 6.821 23.821 23.813 It is worth noting that the slope height H has relatively insignificant
5 10.0 6.727 6.806 6.164 6.337 22.682 22.682 effect on the application of the LSM, as shown in Fig. 9. As seen in the
figure, a peak value of the deviation can be found distinctly for smaller
value of m1 and the maximum deviation occurs in a slope with a height
mechanism, the geometrical variables should satisfy the following of about 8 m. When m1 > 20, the peak value of the deviation can
constrains: β′ ≤ β, 0° < θ0 < θh < 180°, 0° < θ0 < θB < θC hardly be noticed and the deviation with different slope heights can
< θh < 180° and 0 < r′0/r0 < 1.0. nearly be neglected (Dm1 < 0.15%). This indicates that, for a gentle
For quantitatively evaluating the effect of layer numbers m1 and m2 soil slope, the value of H has relatively insignificant effect on the usage
on the calculated results, a deviation Dm1 of the safety factor is in- of the LSM when the soil mass is divided into 20 layers or more in
troduced, which is given as follows practice. This conclusion is similar to that drawn by the authors re-
Fs |m1= 100 Fs |m1 cently [42]. The effect of important parameters on the application of
Dm1 = × 100% the LSM in slope stability analyses was systematically studied by the
Fs |m1= 100 (21)
authors [41,42] recently, where both saturated and unsaturated con-
where Fs |m1= 100 is the safety factor of an unsaturated soil slope with the ditions are considered. It can be found that, the effect of layer number
layer number m1 = 100 and is selected as a benchmark, Fs |m1 is a safety on the usage of the LSM varies with soil types and becomes pronounced
factor of the slope with layer number m1. Note that, the value of m2 is as the suction increases. For clayey soil slope that is investigated in this
determined for a given value of m1 according to Eq. (13). paper, the effect of suction on the slope stability is most pronounced.
This means the aforementioned conclusion is valid for most soils.
5. Validity and comparison
5.3. Effect of suction
5.1. Validity of the optimization procedure
Fig. 10 shows the effect of suction on the seismic stability of un-
The validity of the optimization procedure is checked by comparing saturated clayey soil slopes (n = 2.0, α = 0.005 kPa−1, Sr = 0.16)
the calculated results with published solutions [10] under no suction under different infiltration conditions. The parameters H = 10 m,
conditions. To compare directly, the stability factors are listed in β = 30°, B/H = 2.0, z0 = 0.5H and m1 = 40 are used. It can be seen
Table 2. It can be seen from Table 2 that, for most cases, the stability from the figure that the stability of unsaturated soil slope is reduced
factors presented in this study agree well with Gao et al. [10]. The gradually as the infiltration rate and the seismic intensity increases. The
agreement implies that the optimization procedure is reasonable and slope stability decreases continually with increasing the infiltration rate
effective in finding the critical values of the upper bound solutions. and equals to that of no suction case when q/ks = −1.0. The

4 1.6
Deviation of the safety factor, Dm1 (%)

() ()
Deviation of the safety factor, Dm1 (%)

0 20
5 25
3 10 1.2 30
15 35

2 0.8

1 0.4

0 0.0
5 10 50 100 5 10 50 100
Layer number, m1 Layer number, m1

(a) Effect of ( = 30 ) (b) Effect of ( =5 )


Fig. 8. Effects of parameters (a) internal friction angle ϕ′ (β = 30°), and (b) inclined angle β (ϕ′ = 5°) on application of LSM for gentle unsaturated clayey soil slopes
(H = 10 m, B/H = 2.0, z0 = 0.5H, q = 0 m/s and khE = 0).

303
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

Table 3
Safety factors obtained using LSM for clayey soil slopes under no-flow condition (c′ = 20 kPa, z0 = 0.5H, q = 0 m/s, β = 30°, H = 10 m, B/H = 2.0 and khE = 0.0).
m1 ϕ′ = 0° ϕ′ = 5° ϕ′ = 10° ϕ′ = 15°

Fs Dm1 (%) Fs Dm1 (%) Fs Dm1 (%) Fs Dm1 (%)

4 0.8679 3.72 1.6157 1.19 2.9711 0.58 5.7674 0.23


7 0.8850 1.82 1.6254 0.60 2.9783 0.34 5.7732 0.13
10 0.8908 1.18 1.6289 0.39 2.9809 0.25 5.7758 0.09
15 0.8950 0.71 1.6312 0.24 2.9832 0.17 5.7772 0.06
20 0.8969 0.50 1.6323 0.18 2.9844 0.13 5.7780 0.05
30 0.8988 0.29 1.6334 0.11 2.9857 0.09 5.7788 0.03
60 0.9007 0.08 1.6346 0.04 2.9874 0.03 5.7795 0.02
100 0.9014 0.00 1.6352 0.00 2.9884 0.00 5.7808 0.00

continuous and smooth transition for a soil going from an unsaturated 1.2
state to a saturated state indicates to some extent the validity of the

Deviation of the safety factor, Dm1 (%)


analytical framework. The influence of infiltration on the slope stability
can be easily interpreted in Fig. 3. An increase in the infiltration rate m1 = 5
0.9
leads to an increase in unsaturated soil weight and a decrease in suc-
tion, which all reduce the slope stability. Note that, the effect of suction
on the failure surface varies with soil types [41,42]. When the suction
in soils is large, the failure surface becomes deeper, but the width of 0.6
landslides becomes smaller.

6. Examples
0.3 10
Parametric analyses about the effects of soil type and unsaturated
flow condition on the slope stability have already been studied by many 20
researchers (e.g., [20,40]). In this section, only examples with real soil 0.0 100
parameters are given to demonstrate the 3D effect and the seismic effect 2 4 6 8 10 12 14 16 18 20
on the stability of unsaturated soil slopes. The parameters for different
Slope height, H (m)
soils along wetting paths are given in Table 1.
Fig. 11 shows the change in the seismic stability of unsaturated soil Fig. 9. Deviation versus slope height relation for gentle unsaturated clayey soil
slopes with different constrains of the width-to-height ratio B/H for slopes (β = 30°, B/H = 2.0, ϕ′ = 5°, z0 = 0.5H, q = 0 m/s and khE = 0).
each soil. The parameters H = 10 m, β = 45°, z0 = 0.5H, q = 0 m/s and
m1 = 40 are used. For comparison, the cases considering and ignoring no suction case. When the value of B/H is increased to 10, the corre-
the suction are plotted with “WS” and “NS” in the figure, respectively. sponding safety factors are 2.420 and 1.317, respectively, which both
As expected, the slope stability generally decreases as the seismic in- approach those of the 2D cases (the safety factors are 2.322 and 1.267,
tensity increases. Meanwhile, the 2D stability approach gives a con- respectively). For Soil B (Fig. 11b), the 3D effect and the suction-in-
servative estimation of the slope safety. duced effect on the slope stability are similar with Soil A. However, the
To make a realistic estimation of the slope safety, the 3D effect and safety factors for no suction cases are almost less than 1.0, which are
the suction-induced effect are essential to be considered in the slope much smaller than those considering the suction-induced effect on the
stability analyses. The 3D effect of unsaturated soil slopes is nearly slope stability. The calculated results indicate that ignoring the 3D ef-
identical for different soils and is similar to that of no suction cases fect and the suction-induced effect on the slope stability may lead to
[27]. Conversely, the suction-induced effect varies with soils, is most errors.
pronounced for Soil A and Soil C and decreases in Soil B and Soil D, as
shown in Fig. 11. For Soil A with B/H = 2.0 and khE = 0.3 (Fig. 11a),
the safety factor is 2.933 when the suction is considered but is 1.578 for

Table 4
Safety factors obtained using LSM for clayey soil slopes under no-flow condition (c′ = 20 kPa, z0 = 0.5H, q = 0 m/s, β = 30°, B/H = 2.0, H = 10 m, ϕ′ = 5° and
khE = 0.0).
m1 β = 20° β = 25° β = 30° β = 35°

Fs Dm1 (%) Fs Dm1 (%) Fs Dm1 (%) Fs Dm1 (%)

4 2.2777 0.54 1.8595 0.80 1.6157 1.19 1.4554 1.40


7 2.2849 0.22 1.8671 0.39 1.6254 0.60 1.4657 0.70
10 2.2875 0.11 1.8702 0.23 1.6289 0.39 1.4692 0.46
15 2.2891 0.04 1.8722 0.12 1.6312 0.24 1.4719 0.28
20 2.2895 0.02 1.8730 0.08 1.6323 0.18 1.4731 0.20
30 2.2900 0.00 1.8737 0.04 1.6334 0.11 1.4743 0.12
60 2.2900 0.00 1.8742 0.01 1.6346 0.04 1.4755 0.04
100 2.2900 0.00 1.8745 0.00 1.6352 0.00 1.4761 0.00

304
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

1.8
q/ks
approach of limit analysis theorem. For this, the extended Mohr-
Coulomb failure criterion was introduced to describe the dependency of
0
1.6 - 0.2 shear strength in unsaturated soils on matric suction. Meanwhile, the
- 0.4 3D rotational failure mechanism [27] used for no suction case is ex-
- 0.6 tended to an unsaturated condition. The widely used pseudo-static
Safety factor, Fs

1.4
- 0.8 approach was adopted and the external work rate done by the seismic
- 1.0 (NS) load was incorporated in the energy balance equation to account for the
1.2
effect of seismic excitation on the slope stability. To deal with the
highly variable nature of suction, a semi-analytical method (including
1.0 the SM and the LSM) is developed. By comparing with the existing 3D
solutions, the validity of the analytical framework is demonstrated.
0.8 Based on the calculated results, it can be found that the LSM is ef-
fective in assessing the stability of gentle unsaturated soil slopes and the
0.6 slope height has insignificant effect on the application of the LSM. For a
0.00 0.05 0.10 0.15 0.20 0.25 0.30 gentle soil slope, the LSM is reasonable and effective (Dm1 < 0.15%)
Effective horizontal seismic coefficient, khE when the soil mass is divided into 20 layers or more in practice.
Accounting for the 3D effect and the suction-induced effect can both
Fig. 10. Stability of gentle unsaturated clayey soil slopes under different in- lead to more realistic results. The 3D effect is nearly identical for dif-
filtration conditions (H = 10 m, β = 30°, B/H = 2.0, z0 = 0.5H and m1 = 40) ferent soils while the suction-induced effect varies with soil type. When
(NS represents no suction case). subjected to seismic load, the slope stability is significantly reduced as
the seismic intensity increases. The described framework allows the
7. Conclusions user to quantify the effect of changes in moisture content on effective
stress distribution and effective unit weight distribution in the slope and
This paper develops an analytical framework for evaluating the 3D thus, changes in the slope stability.
seismic stability of gentle unsaturated soil slopes using the kinematical

6 6
WS NS
B/H B/H
5 5 1.0 1.0
2.0 2.0
5.0 5.0
Safety factor, Fs

Safety factor,Fs

4 4 10.0 10.0
2D 2D
3 3

2 2

1 1

0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.00 0.05 0.10 0.15 0.20 0.25 0.30
Effective horizontal seismic coefficient, khE Effective horizontal seismic coefficient, khE
(a) Soil A (b) Soil B

6 6

5 5
Safety factor, Fs

Safety factor,Fs

4 4

3 3

2 2

1 1

0 0
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.00 0.05 0.10 0.15 0.20 0.25 0.30
Effective horizontal seismic coefficient, khE Effective horizontal seismic coefficient , khE
(c) Soil C (d) Soil D
Fig. 11. Safety factor versus horizontal seismic coefficient relation for gentle unsaturated soil slopes (H = 10 m, β = 45°, z0 = 0.5H, q = 0 m/s and m1 = 40): (a) Soil
A; (b) Soil B; (c) Soil C; and (d) Soil D (WS and NS represent with suction and no suction cases, respectively).

305
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

Acknowledgement Public Welfare Technology Research Projects of Zhejiang Province (No.


LGG18D020001) and the China Scholarships Council (No.
This research was financially supported by the National Natural 201806890062) of China.
Science Foundation of China (Nos. 11672172 and 41630633), the

Appendix

The variables z1, z2 and z3 in Eq. (11) are the vertical distances from the middle point of line segments mn¯ , m¯n and m ¯n to the slope toe
elevation, respectively, which can be derived from the geometric relationship in Fig. 6(a) and are given as follows:
1
z1 = z 0 + r0 {2e ( h 0 ) tan
sin e( 0 ) tan
sin sin 0}
2
h
(A1)

1 sin( h + )
z2 = z 0 + r0 {e ( h 0) tan
sin h e( 0 ) tan
sin + e( h 0 ) tan
sin h sin
2 sin( + ) (A2)
1
z3 = z 0 + r0 {e ( h 0) tan
sin e( 0 ) tan
sin }
2
h
(A3)
The variables xc1, xc2 and xc3 in Eq. (12) are given as
4R [1 (a R)2]3 2
x c1 =
3[2 arccos(a R) 2(a R) 1 (a R ) 2 ] (A4)

4R [1 (d R)2]3 2
x c2 =
3[2 arccos(d R) 2(d R) 1 (d R ) 2 ] (A5)

4R [1 (e R)2]3 2
x c3 =
3[2 arccos(e R) 2(e R) 1 (e R ) 2 ] (A6)
The variables z4, z5 and z6 in Eq. (12) are the vertical distances from the center points of areas A(θ) to the slope toe elevation, respectively, which
can be derived from the geometric relationship in Fig. 6(b) and are given as follows:
z7 z9
z 4 = z7 (x c1 a)
R a (A7)
z8 z9
z5 = z 8 (x c2 d)
R d (A8)
z0 z9
z6 = z0 (x c3 e)
R e (A9)
where z7 is the slope height, z8 is the vertical distance from a generic point (point n′) on the slope surface to the slope toe elevation, and z9 is the
vertical distance from a generic point (point m, m′ or m′′) on the slip surface to the slope toe elevation. The variables z7, z8 and z9 can be derived from
the geometric relationship in Fig. 6(a) and are given as follows:

z7 = z 0 + r0 e ( h 0 ) tan
sin h r0 sin 0 (A10)

sin( h + )
z 8 = z 0 + r0 e ( h 0 ) tan
sin h r0 e ( h 0 ) tan
sin
sin( + ) (A11)

z 9 = z 0 + r0 e ( h 0 ) tan
sin h r0 e ( 0 ) tan
sin (A12)
Expressions of the variables aj, θj′, θj, ek, θk′ and θk can be derived from the geometric relationship in Fig. 7, which are given as follows:
r0 sin 0 + (1 j m1 ) H
aj = rm
sin (A13)

r0 e ( j 0 ) tan
sin j (1 j m1 ) H = r0 sin 0 (A14)

sin j sin
j = arctan
cos j sin [sin( j + ) e( h j ) tan sin( h + )] (A15)

r0 sin 0 + H + (n k ) h2
ek = rm
sin (A16)

r0 e ( k 0 ) tan
sin k = r0 sin 0 + [H + (n k ) h2] (A17)

r0 e ( k 0 ) tan
sin k = r0 e ( h 0) tan
sin h + (n k ) h2 (A18)

306
L. Wang, et al. Computers and Geotechnics 110 (2019) 296–307

References [26] Kramer S. Geotechnical earthquake engineering. Prentice Hall; 1996.


[27] Michalowski RL, Drescher A. Three-dimensional stability of slopes and excavations.
Géotechnique 2009;59(10):839–50. https://doi.org/10.1680/geot.8.P.136.
[1] Andrianopoulos KI, Papadimitriou AG, Bouckovalas GD, Karamitros DK. Insight into [28] Michalowski RL, Martel T. Stability charts for 3D failures of steep slopes subjected
the seismic response of earth dams with an emphasis on seismic coefficient esti- to seismic excitation. J Geotech Geoenviron Eng 2011;137(2):183–9. https://doi.
mation. Comput Geotech 2014;55(1):195–210. https://doi.org/10.1016/j. org/10.1061/(ASCE)GT.1943-5606.0000412.
compgeo.2013.09.005. [29] Nadukuru SS, Michalowski RL. Three-dimensional displacement analysis of slopes
[2] Baker R, Shukha R, Operstein V, Frydman S. Stability charts for pseudo-static slope subjected to seismic loads. Can Geotech J 2013;50(6):650–61. https://doi.org/10.
stability analysis. Soil Dyn Earthquake Eng 2006;26(9):813–23. https://doi.org/10. 1139/ cgj-2012-0223.
1016/j.soildyn.2006.01.023. [30] Pan QJ, Xu JS, Dias D. Three-dimensional stability of a slope subjected to seepage
[3] Chen WF. Limit analysis and soil plasticity. Amsterdam: Elsevier; 1975. forces. Int J Geomech 2017;17(8):04017035. https://doi.org/10.1061/(ASCE)GM.
[4] Cornforth DH. Landslides in practice: Investigation, analysis, and remedial/pre- 1943-5622.0000913.
ventative options in soils. Hoboken, N. J.: Wiley; 2005. [31] Papadimitriou AG, Bouckovalas GD, Andrianopoulos KI. Methodology for esti-
[5] Drescher A. Limit plasticity approach to piping in bins. J Appl Mech mating seismic coefficients for performance-based design of earthdams and tall
1983;50(3):549–53. https://doi.org/10.1115/1.3167089. embankments. Soil Dyn Earthquake Eng 2014;56(1):57–73. https://doi.org/10.
[6] Drucker DC, Prager W. Soil mechanics and plastic analysis or limit design. Q Appl 1016/j.soildyn.2013.10.006.
Math 1952;10(2):157–65. https://doi.org/10.1090/qam/48291. [32] Prakash K, Sridharan A, Thejas HK, Swaroop HM. A simplified approach of de-
[7] Duncan JM. State of the art: Limit equilibrium and finite element analysis of slopes. termining the specific gravity of soil solids. Geotech Geol Eng 2012;30(4):1063–7.
J Geotech Eng 1996;122(7):577–96. https://doi.org/10.1061/(ASCE)0733- https://doi.org/10.1007/s10706-012-9521-6.
9410(1996) 122:7(577). [33] Robinson JD, Vahedifard F, Aghakouchak A. Rainfall-triggered slope instabilities
[8] Efremidis G, Avlonitis M, Konstantinidis A, Aifantis EC. A statistical study of pre- under a changing climate: comparative study using historical and projected pre-
cursor activity in earthquake-induced landslides. Comput Geotech 2017;81:137–42. cipitation extremes. Can Geotech J 2017;54:117–27. https://doi.org/10.1139/cgj-
https://doi.org/10.1016/j. compgeo.2016.08.010. 2015-0602.
[9] Gao YF, Zhang F, Lei GH, Li DY, Wu YX, Zhang N. Stability charts for 3D failures of [34] Saada Z, Maghous S, Garnier D. Stability analysis of rock slopes subjected to see-
homogeneous slopes. J Geotech Geoenviron Eng 2013;139(9):1528–38. https://doi. page forces using the modified Hoek-Brown criterion. Int J Rock Mech Min Sci
org/10.1061/(ASCE)GT.1943-5606.0000866. 2012;55(55):45–54. https://doi.org/10.1016/j.ijrmms.2012.06.010.
[10] Gao YF, Zhang F, Lei GH, Li DY. An extended limit analysis of three-dimensional [35] Sloan SW. Geotechnical stability analysis. Géotechnique 2013;63(7):531–72.
slope stability. Géotechnique 2013;63(6):518–24. https://doi.org/10.1680/geot. https://doi.org/10.1680/geot. 12.RL.001.
12.T.004. [36] Swan CC, Seo Y. Limit state analysis of earthen slopes using dual continuum/FEM
[11] Gardner WR. Some steady-state solutions of the unsaturated moisture flow equation approaches. Int J Numer Anal Meth Geomech 1999;23(12):1359–71. https://doi.
with application to evaporation from a water table. Soil Sci 1958;85(4):228–32. org/10.1002/(SICI)1096-9853(199910)23:123.0.CO;2-Y.
https://doi.org/10.1097/00010694- 195804000-00006. [37] Terzaghi K. Mechanism of landslides. Application of geology to engineering practice
[12] Gavin K, Xue J. Design charts for the stability analysis of unsaturated soil slopes. (Berkeley Volume). Washington, DC, USA: Geological Society of America; 1950. p.
Geotech Geol Eng 2010;28(1):79–90. https://doi.org/10.1007/s10706-009-9282-z. 83–123.
[13] Gazetas G, Garini E, Anastasopoulos I, Georgarakos T. Effects of near-fault ground [38] Van Genuchten MT. A closed-form equation predicting the hydraulic conductivity
shaking on sliding systems. J Geotech Geoenviron Eng 2009;135(181):1906–21. of unsaturated soils. Soil Sci Soc Am J 1980;44(5):892–8. https://doi.org/10.2136/
https://doi.org/10.1061/(ASCE)GT.1943-5606.0000174. sssaj1980. 03615995004400050002x.
[14] Gens A, Hutchinson JN, Cavounidis S. Three-dimensional analysis of slides in co- [39] Vahedifard F, Leshchinsky D, Meehan CL. Relationship between the seismic coef-
hesive soils. Géotechnique 1988;38(38):1–23. https://doi.org/10.1680/geot.1988. ficient and the unfactored geosynthetic force in reinforced earth structures. J
38.1.1. Geotech Geoenviron Eng 2012;138(10):1209–21. https://doi.org/10.1061/
[15] Godt JW, Baum RL, Lu N. Landsliding in partially saturated materials. Geophys Res (ASCE)GT.1943-5606.0000701.
Lett 2009;36(2):206–18. https://doi.org/10.1029/2008GL035996. [40] Vahedifard F, Leshchinsky D, Mortezaei K, Lu N. Effective stress-based limit-equi-
[16] Griffiths DV, Lu N. Unsaturated slope stability analysis with steady infiltration or librium analysis for homogeneous unsaturated slopes. Int J Geomech
evaporation using elasto-plastic finite elements. Int J Numer Anal Meth Geomech 2016;16(6):04016003. https://doi.org/10.1061/(ASCE)GM.1943-5622.0000554.
2005;29(3):249–67. https://doi.org/10.1002/nag.413. [41] Wang L, Hu W, Sun D, Li L. 3D stability of unsaturated soil slopes with tension
[17] Griffiths DV, Marquez RM. Three-dimensional slope stability analysis by elasto- cracks under steady infiltrations. Int J Numer Anal Meth Geomech
plastic finite elements. Géotechnique 2007;57(6):537–46. https://doi.org/10.1680/ 2019;43(6):982–1004. https://doi.org/10.1002/nag.2889.
geot.2007.57.6.537. [42] Wang L, Sun D, Li L. 3D stability of partially saturated soil slopes after rapid
[18] Le TMH, Gallipoli D, Sánchez M, Wheeler S. Stability and failure mass of un- drawdown by a new layer-wise summation method. Landslides
saturated heterogeneous slopes. Can Geotech J 2015;52:1747–61. https://doi.org/ 2019;16(2):295–313. https://doi.org/10.1007/s10346-018-1081-2.
10.1139/cgj-2014-0190. [43] Xu JS, Yang XL. Effects of seismic force and pore water pressure on three dimen-
[19] Li AJ, Merifield RS, Lyamin AV. Three-dimensional stability charts for slopes based sional slope stability in nonhomogeneous and anisotropic soil. KSCE J Civ Eng
on limit analysis methods. Can Geotech J 2010;47(12):1316–34. https://doi.org/ 2018;22(5):1720–9. https://doi.org/10.1007/s12205-017-1958-y.
10.1139/T10-030. [44] Yeh TCJ. One-dimensional steady-state infiltration in heterogeneous soils. Water
[20] Li ZW, Yang XL. Stability of 3D slope under steady unsaturated flow condition. Eng Resour Res 1989;25(10):2149–58. https://doi.org/10.1029/WR025i010p02149.
Geol 2018;242:150–9. https://doi.org/10.1016/j.enggeo.2018.06.004. [45] Zhang F, Gao YF, Wu YX, Zhang N, Qiu Y. Effects of vertical seismic acceleration on
[21] Likos WJ, Lu N, Godt JW. Hysteresis and uncertainty in soil water-retention curve 3D slope stability. Earthquake Eng Eng Vibration 2016;15(3):487–94. https://doi.
parameters. J Geotech Geoenviron Eng 2014;140(4):04013050. https://doi.org/10. org/10.1007/s11803 -016-0338-9.
1061/(ASCE)GT.1943-5606.0001071. [46] Zhao LH, Cheng X, Zhang Y, Li L, Li DJ. Stability analysis of seismic slopes with
[22] Lu N, Godt JW, Wu DT. A closed-form equation for effective stress in unsaturated cracks. Comput Geotech 2016;77:77–90. https://doi.org/10.1016/j.compgeo.2016.
soil. Water Resour Res 2010;46(5):W05515. https://doi.org/10.1029/ 04.007.
2009WR008646. [47] Zhou XP, Cheng H. Stability analysis of three-dimensional seismic landslides using
[23] Lu N, Likos WJ. Unsaturated soil mechanics. New York: John Wiley & Sons; 2004. the rigorous limit equilibrium method. Eng Geol 2014;174(8):87–102. https://doi.
[24] Lyamin AV, Sloan SW. Upper bound limit analysis using linear finite elements and org/10.1016/j.enggeo. 2014.03.009.
non-linear programming. Int J Numer Anal Meth Geomech 2002;26(2):61–77. [48] Zienkiewicz OC, Humpheson C, Lewis RW. Associated and non-associated visco-
https://doi.org/10.1002/nag.198. plasticity and plasticity in soil mechanics. Géotechnique 1975;25(4):671–89.
[25] Jing L, Liang H, Li Y, Liu C. Characteristics and factors that influenced damage to https://doi.org/10.1680/geot.1975.25.4.671.
dams in the Ms 8.0 Wenchuan earthquake. Earthquake Eng Eng Vibration
2011;10(3):349–58. https://doi.org/10.1007/s11803-011-0071-3.

307

Das könnte Ihnen auch gefallen