Sie sind auf Seite 1von 15

574

Three-dimensional strength-reduction finite


element analysis of slopes: geometric effects
T.-K. Nian, R.-Q. Huang, S.-S. Wan, and G.-Q. Chen

Abstract: The vast majority of slopes, both natural and constructed, exhibit a complex geometric configuration and three-di-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

mensional (3D) state, whereas slopes satisfying the assumption of plane strain (infinite length) are seldom encountered. Ex-
isting research mainly emphasizes the 3D dimensions and boundary effect in slope stability analysis; however, the effect of
complex geometric ground configuration on 3D slope stability is rarely reported. In this paper, an elastoplastic finite-element
method using strength-reduction techniques is used to analyze the stability of special 3D geometric slopes. A typical 3D
slope underlain by a weak layer with groundwater is described to validate the numerical modeling, safety factor values, and
critical slip surface for the 3D slope. Furthermore, a series of special 3D slopes with various geometric configurations are
analyzed numerically, and the effects of turning corners, slope gradient, turning arcs, and convex- and concave-shaped sur-
face geometry on the stability and failure characteristics of slopes under various boundary conditions are discussed in detail.
Key words: slope stability, three-dimensional, strength-reduction finite-element method, geometric configuration, safety fac-
tor, critical slip surface.
Résumé : La grande majorité des pentes, autant naturelles que construites, comportent une configuration géométrique com-
plexe et un état tridimensionnel (3D), tandis que les pentes répondant au critère de déformation en plan (longueur infinie)
sont rarement rencontrées. La recherche existante se penche surtout sur les trois dimensions et les effets de frontière dans
les analyses de stabilité de pente; cependant, l’effet de la configuration géométrique complexe du sol dans l’analyse de stabi-
For personal use only.

lité de pente en 3D est rarement discuté. Dans cet article, une méthode par éléments finis élasto-plastique utilisant des tech-
niques de réduction de la résistance est d’abord utilisée pour analyser la stabilité de pentes à géométrie 3D spéciales. Une
pente 3D typique placée sur une couche de sol à faible résistance avec une nappe phréatique est décrite pour valider le mo-
dèle numérique, les valeurs du facteur de sécurité et la surface critique de glissement de la pente en 3D. De plus, une série
de pentes 3D spéciales avec des configurations géométriques variées sont analysées numériquement. Les effets de certains
paramètres, dont les coins tournants, les gradients de pente, les arcs tournants et les surfaces avec des géométries convexes
et concaves, sur la stabilité et sur les caractéristiques de rupture des pentes dans des conditions variées sont discutés en dé-
tails.
Mots‐clés : stabilité de pente, tridimensionnel, méthode par éléments finis de réduction de la résistance, configuration géo-
métrique, facteur de sécurité, surface critique de glissement.
[Traduit par la Rédaction]

Introduction sult obtained is very conservative (Duncan 1996a; Griffiths


and Marquez 2007). On the whole, the results for 2D analysis
In slope stability analysis, two-dimensional (2D) plane are more conservative from the viewpoint of engineering
strain analysis techniques are commonly used for simplicity safety, and moreover the actual stability and geometry cannot
and wide applicability. All slope failures are, however, three- be appropriately considered without the third dimension.
dimensional (3D) in nature, especially in the cases of natural Therefore, a 3D slope stability analysis is required to obtain
complex geometric slopes or overloaded slopes (Wei et al. more reasonable solutions in some cases. However, the re-
2009). According to previous studies (Chen and Chameau sults from 3D stability analysis are not always more reliable
1985; Leshchinsky and Baker 1986; Cavounidis 1987; Hungr than 2D results; under some special combinations of soil and
1987; Duncan 1996a; Chugh 2003; Griffiths and Marquez groundwater properties, complex dimensions, boundary con-
2007), the safety factor for a 2D slope is generally less than ditions, and complex geometries (e.g., ridges and corners),
that for a 3D slope. In particular, when the most pessimistic the reverse conclusion can probably be drawn (Chen and
section in the 3D problem is selected for 2D analysis, the re- Chameau 1985; Hungr 1987; Bromhead and Martin 2004;
Received 18 April 2011. Accepted 1 February 2012. Published at www.nrcresearchpress.com/cgj on 27 April 2012.
T.-K. Nian and S.-S. Wan. Institute of Geotechnical Engineering, State Key Laboratory of Coastal and Offshore Engineering, School of
Civil and Hydraulic Engineering, Dalian University of Technology, Dalian 116024, China.
R.-Q. Huang. State Key Laboratory of Geohazard Prevention and Geoenvironmental Protection, Chengdu University of Technology,
Chengdu 610059, China.
G.-Q. Chen. Department of Civil and Structural Engineering, Faculty of Engineering, Kyushu University, Fukuoka, 812-8581, Japan.
Corresponding author: Ting-Kai Nian (e-mail: tknian@dlut.edu.cn).

Can. Geotech. J. 49: 574–588 (2012) doi:10.1139/T2012-014 Published by NRC Research Press
Nian et al. 575

Sainak 2004; Vaughan 2008). In these cases, the slope di- Fig. 1. Relationship between SRF and dimensionless displacement.
mensions and real-world geometric configuration cannot be
disregarded, and a more reasonable 3D slope stability analy-
sis method should be used to solve the problem. Otherwise,
the analysis will lead to wasted engineering construction ef-
fort or an incorrect evaluation of potential landslide mecha-
nisms.
Currently, the 3D limit equilibrium method (LEM) (Dun-
can 1996b; Stark and Eid 1998; Arellano and Stark 2000;
Huang and Tsai 2000; Huang et al. 2002; Jiang and Yama-
gami 2004; Cheng et al. 2007; Cheng and Yip 2007) and the
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

finite-element method (FEM) or finite-difference method


(FDM) (Ugai and Leshchinsky 1995; Cai et al. 1998; Jeremic
2000; Chugh 2003; Sainak 2004; Griffiths and Marquez tanf0 c0
2007; Wei et al. 2009) with a strength-reduction technique ½1 tanfR0 ¼ ; cR0 ¼
SRF SRF
are widely popular for analyzing the stability of a 3D slope,
even though these approaches are more time-consuming than are used in the analysis, where SRF is a “strength-reduction
others. The major limitation in 3D LEM is the lack of a suit- factor” and c′ and f 0 are the original shear strength para-
able method for locating the critical general 3D slip surface meters. To obtain a true safety factor, the strength-reduction
(Cheng et al. 2005; Wei et al. 2009). However, 3D FEM factor needs to be increased gradually until the reduced
with a strength-reduction technique can simultaneously pro- strength parameters cR0 and fR0 bring the slope to a limit equi-
vide the safety factor and the critical slip surface (shape and librium state. In this process, the factor SRF is assumed to
location), together with the stress, deformation, and progres- apply equally to both c′ and tanf 0 . When the slope reaches
sive shear failure of the slope, and it has been widely and the limit failure state, the safety factor (FOS) of the slope is
successfully used in the engineering field with the develop- equal to the strength-reduction factor, and FOS = SRF. This
ment of computer technology. Moreover, 3D analysis can ex- definition of FOS is identical to that used in the limit equili-
For personal use only.

actly reflect the slope dimensions, boundary conditions, and brium methods and is widely used by a large number of
a realistically complex geometric configuration in three x–y–z authors (Zienkiewicz et al. 1975; Ugai and Leshchinsky
directions. In this way, the computational accuracy can be 1995; Duncan 1996b; Dawson et al. 1999; Griffiths and
greatly improved, and the fundamental nature of slope fail- Lane 1999; Zheng et al. 2006; Griffiths and Marquez 2007;
ure mechanisms can be fully revealed. Wei et al. 2009). The safety factor and critical slip surface
However, existing FEM studies have mainly emphasized obtained are also entirely identical to those obtained from
3D effects and boundary conditions in slope stability analysis LEM (Manzari and Nour 2000). The overall procedure can
be performed by having the program solve the problem re-
(Chugh 2003; Griffiths and Marquez 2007; Wei et al. 2009),
peatedly using a sequence of user-specified SRF values.
whereas the effect of the geometric ground configuration on
3D slope stability is rarely reported (Sainak 2004; Griffiths
Slope failure
and Marquez 2007). Based on this observation, a 3D
strength-reduction FEM using an elastic, perfectly plastic When the algorithm cannot converge within a user-specified
maximum number of iterations (typically set to 1000), the
model is used here to investigate the stability of special 3D
implication being that no stress distribution can be found
geometric slopes. A typical 3D slope underlain by a weak
that is simultaneously able to satisfy both the failure crite-
layer with groundwater is used as an illustration to validate
rion and the overall equilibrium criterion, failure is said to
the numerical modeling approach. Furthermore, a series of have occurred. Slope failure and numerical nonconvergence
special 3D slopes with various geometric ground configura- occur simultaneously and are accompanied by a dramatic
tions are numerically analyzed, and the effects of turning cor- increase in the nodal displacements within the mesh (Grif-
ners, slope gradient, turning arcs, and convex- and concave- fiths and Lane 1999; Griffiths and Marquez 2007). At that
shaped surface geometry on the stability and failure charac- time, the safety factor of the slope is equal to the strength-
teristics of slopes under various boundary conditions are dis- reduction factor, and FOS = SRF. Figure 1 shows a typical
cussed in detail. graph of SRF against Edmax/gH2 (a dimensionless displace-
ment), where E is the Young’s modulus of the soil, dmax is
Three-dimensional FEM by strength- the maximum nodal displacement component at conver-
reduction technique gence, g is the unit weight of the soil, and H is the slope
height.
Strength-reduction technique
In traditional slope stability analysis, the safety factor is Numerical execution
defined as the ratio of the average shear strength of the soil The finite element (FE) program, Abaqus (Hibbitt, Karls-
to the average shear stress developed along the critical sliding son, and Sorensen, Inc. 2005), has been improved in certain
surface. Based on this definition, the current approach is to aspects to carry out the computation. Three-dimensional anal-
use a shear strength-reduction technique (Griffiths and Mar- ysis of elastic, perfectly plastic soils with a Mohr–Coulomb
quez 2007) in which reduced shear strength parameters cR0 failure criterion assuming zero dilation is performed, and
and fR0 , given by no tensile strength is considered in the present paper.

Published by NRC Research Press


576 Can. Geotech. J. Vol. 49, 2012

Fig. 2. 3D slope conceptual model and boundary conditions. (a) 3D slope sketch; (b) side view (x–y plane); (c) plan view (x–z plane).

Although any 3D FE algorithm could be used in principle, • A slope that is composed of a strong soil near the abutments
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

the current work uses the popular 20-node hexahedral ele- surrounding a weak soil in the central parts of the slope.
ment configuration with reduced integration in the stiffness In other words, the strong soil surrounds the weaker soil.
matrix generation and stress re-distribution phases of the al- • The confinement of the vertical abutments or sloping
gorithm. The nonconvergence criterion accompanied by a abutments of an earth dam or canyon wall.
dramatic nodal displacement increase is chosen to evaluate Type 1, unrestrained BC, is used to represent a contact with a
the overall stability of a slope (Griffiths and Marquez rigid, smooth abutment that can provide a reactive thrust, but
2007). Soil–water coupling and unsaturated, non-steady-state no in-plane shear restraint, and it can be used in the follow-
seepage are not considered in the present analysis. ing three cases:
• Displacement is constrained only in the out-of-plane direc-
Verification tion at the model ends or the two sides of the slope. The
To check the applicability of the 3D FE program with side or end boundary is placed not far enough from the
strength-reduction method, a conceptual 3D slope model and region where slope failure is likely to occur.
the corresponding boundary conditions (BCs) are introduced • The whole slope is composed of the same soil with homo-
in Fig. 2. Figure 2 shows a homogeneous 3D slope in which geneity and isotropy.
• For a laboratory model of a slope built in a wooden or steel
For personal use only.

the geometry and dimensions in the x–y plane are extended


by a distance L/2 (assuming symmetry) in the z direction. container with glass walls, the displacement is only con-
The boundary conditions can be generally classified into strained in the out-of-plane direction at the model ends
three types: or the two sides of the slope.
1. Unrestrained BC (smooth–smooth): Both upper and lower Type 2 as described here implies a symmetric analysis about
sides as depicted in the plane view (Fig. 2c) are restrained in the plane z = L/2, and therefore only half of the actual ex-
the z-direction, but free to move in the x- and y-directions. tended length L of the slope is analyzed for simplification.
The left and right sides as depicted in both views are Type 2 is assumed for the numerical analysis of straight
allowed to move freely in the y- and z-directions, but slopes in example 1 below.
not in the x-direction. In the bottom plane of the model,
all movements are restrained (i.e., ux, y, z = 0, where u is Example 1
the displacement). A typical slope example, including a weak layer and
2. Semi-restrained BC (rough–smooth): The lower side (z = groundwater, is shown in Fig. 3, and its geometric and mate-
L/2) as depicted in the plane view is restrained in the z- rial properties are listed in Table 1. The slope was first con-
direction, but free to move in the x- and y-directions. sidered by Zhang (1988) using an extended Spencer’s
The upper side (z = 0) is completely restrained in the method. Later, the model slope was widely used by various
x-, y-, and z-directions. The other conditions are the investigators as part of the validation of their particular 3D
same as for type 1. slope stability methods (Lam and Fredlund 1993; Huang and
3. Fully restrained BC (rough–rough): Both upper and lower Tsai 2000; Chen et al. 2001, 2003b; Huang et al. 2002; Grif-
sides as depicted in the plane view are restrained in all fiths and Marquez 2007). In Zhang’s study, the four cases of
three x-, y-, and z-directions, and the other conditions are homogeneous soil slope, nonhomogeneous soil slope with a
the same as for type 1. thin weak layer, homogeneous soil slope with a piezometric
line, and nonhomogeneous soil slope with a piezometric line
Among the three types of BC, the fully restrained BC or the
are investigated, and solutions are obtained for their safety
unrestrained BC is the most applicable to a real example, the
factors and the corresponding critical slip surfaces. Moreover,
semi-restrained BC is rarely used in the real world, but is of-
the proposed critical slip surface for the homogeneous slope
ten used in FE computation with a half-model for simplifying
(case 1) is from limit equilibrium considerations a symmetric
the fully restrained BC. ellipsoidal surface, meaning that the slip surface is circular in
In particular, type 3 (fully restrained BC) can be used to the x–y plane, but elliptical in the z-direction.
represent a rigid contact with no possibility of movement (e.g., To make adequate comparisons with the literature (Zhang
Chugh 2003; Griffiths and Marquez 2007). The applicable 1988; Griffiths and Marquez 2007) possible, the dimensions
conditions in practice can be illustrated as of the 3D slope are the same as those given in Griffiths and
• No displacement at the model ends or the two sides of the Marquez (2007). L is equal to 25 m, but the actual mesh ex-
slope. The side–end boundary is placed far enough from tended length varied by half this amount because of symme-
the region where slope failure is likely to occur. try. Both the coarse and fine meshes for the case 1 slope with

Published by NRC Research Press


Nian et al. 577

Fig. 3. Cross section of 3D slope with free surface and weak layer reported by Zhang (1988). c′, cohesion; f 0 , friction angle; g, unit weight;
R, radius.

Case 1 circular R=
slip surface 24
.4 20
m
g=18.8kN/m3
c'=29kPa, f'=20° 15
2

(m)
1
10
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

e
etric lin
Piezom

5
Weak layer
c'=0, f'=10°
0
50 40 30 20 10 0
(m)

Table 1. Material properties of examples 1 and 2.

Example 1
Material properties Upper soil layer Weaker layer Example 2
Friction angle, f′ (°) 20 10 20
Cohesion, c′ (kPa) 29 0 40
For personal use only.

Dilation angle, j (°) 0 0 0


Young’s modulus, E′ (kPa) 1.0 × 104 1.0 × 104 1.0 × 105
Poisson’s ratio, n′ 0.25 0.25 0.30
Unit weight, g (kN/m3) 18.8 19.5 20.0

Fig. 4. Typical fine and coarse 3D slope meshes of 20-node hexahedral elements (L/H = 2): (a) fine mesh; (b) coarse mesh.

L = 2H as indicated in Fig. 4 were run to illustrate the sensi- dary conditions. The extended length in the z-direction (see
tivity of results to mesh refinement. The relationship between Fig. 2) is chosen to be 2H to compare with known publica-
the strength-reduction factor (SRF) and dimensionless dis- tions (Zhang 1988; Griffiths and Marquez 2007).
placement is shown in Fig. 1 for two types of meshes; the
obtained safety factors based on the iterative nonconvergence Safety factor for the four cases
criterion mentioned above are, respectively, FOS = 2.11 and
2.15 for the fine and coarse meshes. Because the difference Case 1: Homogeneous slope
between the FOS values is small, the coarse mesh form is The 3D slope shown in Fig. 4b can be regarded as a sim-
used in the following analysis. ple homogeneous soil slope for which a weak layer and
In the present analysis, the same four cases as in Zhang groundwater are not considered. Three-dimensional strength-
(1988) are used, and assuming model symmetry, the half reduction FEM is used to compute the safety factor and the
slope (L/2) is used to compute the safety factor and the crit- critical slip surface for the slope. The result obtained from
ical slip surface under semi-restrained (rough–smooth) boun- this method is 2.15, which is in good agreement with the

Published by NRC Research Press


578 Can. Geotech. J. Vol. 49, 2012

Table 2. Comparison of slope safety factors for case 1.

Source Factor
Average of 2D methods by Fredlund and Krahn (1977) 2.034
3D LEM by Zhang (1988) 2.122
3D UB-LAM by Chen et al. (2001) 2.262
3D-LEM by Chen et al. (2003b) 2.187
3D-LEM CLARA (software package) by Hungr (1987) 2.167
3D-FEM by Griffiths and Marquez (2007) 2.17
Present solution 2.15
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

Fig. 5. Deformed mesh at failure for four cases. (a) case 1; (b) case 2; (c) case 3; (d) case 4.
For personal use only.

other solutions obtained from different methods and listed in Table 3. Comparison of safety factor values for a 3D slope for
Table 2. The maximum difference is less than 5%. The de- Case 2.
formed mesh at failure (corresponding to the nonconverged
Source Factor
solution when SRF = 2.15), can be shown in Fig. 5a. By in-
spection of the critical slip surfaces shown in Fig. 5a, it can Average of 2D methods by Fredlund and Krahn (1977) 1.383
be seen that both the failure mechanism and the shape and 3D LEM by Zhang (1988) 1.553
location of that surface are somewhat similar to those ob- 3D UB-LAM by Chen et al. (2001) 1.717
tained by 3D LEM and shown in Fig. 3. 3D-LEM by Chen et al. (2003b) 1.603
3D-LEM CLARA (software package) by Hungr (1987) 1.620
Case 2: Nonhomogeneous slope with a thin weak layer 3D-FEM by Griffiths and Marquez (2007) 1.58
A weak layer is assumed to exist in this 3D slope, as Present solution 1.59
shown in Fig. 3. The safety factor of the slope, as calculated
by strength-reduction FEM, is 1.59, which shows good con-
sistency with the other results from different authors listed in Case 3: Homogeneous slope with groundwater
Table 3, except for a slightly greater difference (though still A homogeneous slope with a piezometric line as shown in
less than 10%) between the result from the present method Fig. 3 is considered in the present analysis. The obtained
and that from Chen et al. (2001). safety factor of the slope based on 3D strength-reduction
The deformed FE mesh for the whole slope is shown in FEM is 1.86, which is very close to the 3D solution value of
Fig. 5b. It can be seen that the critical slip surface for the 1.831 obtained by Zhang (1988), while the average of the 2D
3D slope that includes a thin weak layer shows a “dog-leg” methods (including the ordinary method, simplified Bishop
path in the x–y plane and is extremely similar to the surface method, Spencer’s method, Janbu’s simplified method, Jan-
calculated by Griffiths and Marquez (2007). Moreover, the bu’s rigorous method, and the Morgenstern–Price method) as
failure mechanism of the slope that includes a thin weak determined by Fredlund and Krahn (1977) is 1.799. The de-
layer in the foundation is completely different from that de- formed FE mesh is shown in Fig. 5c. Compared with case 1
scribed in case 1 (Fig. 5a). for a homogeneous soil slope, it is apparent that the safety

Published by NRC Research Press


Nian et al. 579

factor decreases dramatically from 2.15 to 1.86 for the 3D

Case 4: nonhomogeneous slope with


slope when groundwater is included, and the critical slip sur-
face is also deeper for the groundwater condition, as can be
seen by comparison of Figs. 5c and 5a. However, it is worth

weak layer and groundwater


mentioning that when further compared with the circular slip
surfaces shown in Fig. 3, the failure mechanism in the
present case is fully similar to that shown in Fig. 3 deter-
mined by 3D-LEM and to that shown in Fig. 5c determined
by 3D-FEM.

–0.6 (32%)

–0.3 (19%)

Shallower

Speedup
Case 4: Nonhomogeneous slope with groundwater

4.97e–02
8.49e–02
Deeper
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

1.298
A nonhomogeneous slope underlain by a thin weak layer
with groundwater as shown in Fig. 3 is investigated in this
case. Three-dimensional strength-reduction FEM is used, and

——————————————————————————————————————————→
————————————————————————————→
the obtained safety factor of the slope is 1.298, which is

Case 3: homogeneous slope with

————————————→
———————————————————————————→
somewhat smaller than the solution (1.441) obtained by

———————————→
Zhang (1988), but much closer to the average of the 2D
methods (1.258) as determined by Fredlund and Krahn
(1977). The maximum difference is still less than 10%. Fig-
ure 5d shows the deformed FE mesh when iteration noncon-
vergence occurs. It can be seen that the failure mechanism

groundwater
and the critical slip surface are fairly similar to case 2.

–0.3 (14%)

7.51e–02
2.16e–01
deeper
Comparison of results for the four cases

1.86
The four cases described above for the slope underlain by
a thin weak layer and groundwater are summarized in Table 4
For personal use only.

to show the differences in safety factor and the critical slip

——————————————————————————→

——————————————————————————→
Case 2: nonhomogeneous slope

surface. The safety factor of the slope is remarkably reduced


when a weak layer or groundwater is present, and the effect
of the weak layer on the safety factor is somewhat larger
Table 4. Comparison of safety factors and critical slip surfaces of the 3D slope for example 1.

than that of groundwater in this example. The critical slip


surface can also be outlined using the deformed mesh and
with weak layer

the equivalent plastic strain zone, and the simplified 2D fail-


ure modes in the x–y plane are presented in Table 4. The crit-
–0.6 (28%)

ical slip surface in the slope containing a weak layer (cases 2


and 4) is evident along the thin weak layer and is completely shallower
1.59

1.25
2.37
different from the surface with groundwater (case 3 or 1).
The zone of high equivalent plastic strain in the slope is also
evident on the 2D plots; slope reinforcement should be em-
phasized in these zones. The maximum equivalent plastic
Case 1: homogeneous slope

——————————→

——————————→

strain value on the slope surface at the time of numerical iter-


ative nonconvergence is also given in Table 4. When the
maximum equivalent plastic strain values for the four cases
are compared, it is clear that iterative nonconvergence tends
to speed up, as shown in Table 4.

Effect of three-dimensional complex


2.15

3.67
3.05

geometry for corner slopes


Numerical iterative nonconvergence

Most natural and constructed slopes exhibit a complex


surface; zones of high equiva-

geometric configuration and 3D state. This is especially clear


Max. equiv. plastic strain, 3pmax
lent plastic strain enclosed in
Shape and location of critical

in instances such as a corner defined by the intersection of


Depth of critical slip surface

two embankments or an arc-shaped road cutting a slope


Difference between cases

around a mountainous area. Engineering practice has shown


FOS (present method)

that geometric configuration has an important impact on


slope stability and therefore should be considered fully in
Speedup order

evaluating the safety factor and critical slip surface for a


Parameters

complex 3D slope. In the following example, a representative


Edmax/gH2

3D complex slope with various geometric configurations,


such as turning corners, slope gradients, and turning arcs, is
discussed in detail.

Published by NRC Research Press


580 Can. Geotech. J. Vol. 49, 2012

Fig. 6. Top views with different corners, side view, and FE mesh for a gentle 3D slope. (a-1) Top view of 90° corner; (a-2) top view of 135°
corner; (a-3); top view of 180° corner (straight slope). (b) Side view. (c) 3D FE model and mesh representation. All dimensions in metres.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13
For personal use only.

Table 5. Safety factor for gentle 3D slopes with different corners and BCs.

Safety factor
Turning corner of slope Unrestrained Semi-restrained Fully restrained
90° 2.690 2.710 2.860
135° 2.655 2.695 2.810
180° (straight slope) 2.590 2.655 2.780
Plane strain solution 2.555 2.555 2.555

Example 2 for the 3D slope — 90°, 135°, and 180° (straight slope) —
A three-dimensional complex geometric slope is illustrated are considered in the present analysis. For each corner case,
in Fig. 6. Top views with different turning corners (90°, three types of constraints — unrestrained, semi-restrained,
135°, and 180°), a side view, and a typical 3D FE model and fully restrained boundary conditions as described ear-
and mesh representation are shown, respectively, in Figs. 6a, lier — are, respectively, incorporated into the three-dimen-
6b, and 6c, in which the 3D slope with 180° turning corner is sional FE analysis using the strength-reduction method. The
completely identical to the common 3D slope without a cor- safety factors of the 3D slope corresponding to various cor-
ner (straight slope). The material properties of the 3D slope ners and constraints are calculated and listed in Table 5.
are listed in Table 1. The height of the slope is H = 10 m It can be seen from Table 5 that the safety factor for a gen-
and c′/gH = 0.2. The slope is inclined at an angle of 26.57 tle 3D slope decreases as the slope turning corner increases
(1V:2H, where V is vertical and H is horizontal) from the from 90° to 180° (straight slope) regardless of the BCs. This
horizontal, and the boundary condition is chosen to be types indicates that the safety factor of a 3D slope with a corner is
1, 2, and 3, respectively, as mentioned earlier. For instance, much larger than for one without a corner, such as the 180°
type 1 is represented by vertical rollers on the left base and (straight slope) case, but the difference is less than 5%. On
right boundary (Fig. 6b) and fully fixed at the base (Figs. 6b, the other hand, the safety factor for a gentle 3D slope will
6c). The constitutive behavior of the soil continuum is mod- increase as the boundary constraints for a given slope corner
eled in the present analysis using an elastic, perfectly plastic are strengthened (from unrestrained to fully restrained BC),
Mohr–Coulomb failure criterion combined with nonassoci- but the change is in the range of less than 10%. Comparing
ated flow laws. The safety factor for the slope, obtained by the effects of various boundary conditions with those of slope
plane strain analysis, is 2.555. corners, it is clear that the effect of the boundary constraints
on the safety factor for a gentle 3D slope is somewhat larger
Effect of a turning corner on the slope than that of a slope corner.
The 3D slopes with turning corners exhibit a symmetric dis- To further investigate the variations in critical slip surface
tribution around the central axis through the corner, and the under various boundary conditions and slope corners, the de-
length of each border is 20 m (Fig. 6). Three turning corners formed FE mesh (Fig. 7) and the horizontal displacement dis-

Published by NRC Research Press


Nian et al. 581

Fig. 7. Deformed FE mesh for a gentle 3D slope under two BCs (90° corner): (a) unrestrained; (b) fully restrained.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

Fig. 8. Horizontal displacement distribution and shear-slip zone contours for a gentle 3D slope under fully restrained BC (90° corner):
(a) horizontal displacement distribution; (b) shear-slip zone contours.
For personal use only.

Fig. 9. Deformed FE meshes for a gentle 3D slope under two BCs (135° corner): (a) unrestrained; (b) fully restrained.

tribution and equivalent plastic strain contours (Fig. 8) are Comparing the above two corner cases (90° and 135°)
shown for a 90° slope corner under two constraints. Similar based on the horizontal displacement distribution and plastic
plots for a 135° slope corner with two constraints can be strain contours, the turning corner acts to some extent as a
found in Figs. 9 and 10, respectively. The results for the constraint to resist potential sliding on a gentle 3D slope
180° straight slope are not reported here because this is a under unrestrained BCs, but the constraint effect is the most
very common 3D slope and has been analyzed by many in- significant in the fully restrained state. In particular, equiva-
vestigators (Zhang 1988; Chen et al. 2001; Chugh 2003; lent plastic strain contours show that the constraint effects
Griffiths and Marquez 2007; Wei et al. 2009). become more noticeable when the turning corner is smaller

Published by NRC Research Press


582 Can. Geotech. J. Vol. 49, 2012

Fig. 10. Contours of shear-slip zone for a gentle 3D slope under fully restrained BC (135° corner).
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13
For personal use only.

Fig. 11. Top views with different corners and side view for a vertical 3D slope. (a) Top view of 90° corner; (b) top view of 135° corner; (c)
side view.

(e.g., a 90° corner angle) regardless of the BCs. This is ex- Effect of the slope gradient
actly why the safety factor for a gentle 3D slope with a 90° In the previous section, the effect of the slope-turning cor-
corner angle is far greater than for the other two corner-turn- ner on slope stability and failure characteristics under various
ing cases (135° and 180°). On the other hand, the horizontal constraints is fully discussed for a gentle 3D slope (1V:2H),
displacement of the slope crest at the corner presents a max- and certain corner effects are found to exist for this gentle
imum value and is in the most dangerous state under fully re- slope. Now, to investigate special phenomena for a steep
strained conditions. In addition, the deformation and failure slope, the gentle slope investigated above will be replaced by
modes under the fully restrained BC are entirely different a vertical 3D slope. The corresponding top views with differ-
from that under unrestrained BC, in which the slope will fail ent corners (90° and 135°) and a side view are shown, re-
along or close to the free boundary. spectively, in Figs. 11a, 11b, and 11c. The three types of

Published by NRC Research Press


Nian et al. 583

Table 6. Safety factor for vertical 3D slopes with different corners and BCs.

Safety factor
Turning corner of slope Unrestrained Semi-restrained Fully restrained
90° 1.172 1.197 1.209
135° 1.205 1.215 1.240
180° (no corner) 1.210 1.235 1.273
Plane strain solution 1.206 1.206 1.206

Fig. 12. Deformed FE meshes for a vertical 3D slope under two BCs (90° corner): (a) unrestrained; (b) fully restrained.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13
For personal use only.

Fig. 13. Horizontal displacement distribution and shear-slip zone contours for a vertical 3D slope under fully restrained BC (90° corner:
(a) horizontal displacement distribution; (b) shear-slip zone contours.

BCs are carried out, and the FE calculation mesh and constit- problem (Table 6), the solution obtained under such a case
utive soil model are identical to those used in previous exam- is not certain to be the minimum one. In other words, the
ples. The calculated results for the safety factors of the safety factor for a vertical 3D slope with a 90° or 135° corner
vertical 3D slope obtained using the FEM are listed in Ta- angle is somewhat smaller than that for a 2D slope. This sit-
ble 6. uation can be commonly encountered in practical engineering
It can be seen from Table 6 that the safety factor for a ver- situations such as the excavation of a deep building pit. For
tical 3D slope increases as the slope corner angle increases, slope engineering, not only the safety factor, but also the de-
irrespective of the BCs. This is entirely opposite to the result formation and the critical slip surface must all be the focus of
obtained for the gentle slope discussed previously. If a 3D further research. The deformed FE mesh (Fig. 12) and hori-
slope with a 180° corner (straight slope) under an unre- zontal displacement distribution and equivalent plastic strain
strained condition is almost equivalent to the plane strain contours (Fig. 13) are shown for a vertical 3D slope with

Published by NRC Research Press


584 Can. Geotech. J. Vol. 49, 2012

Fig. 14. Deformed FE meshes for a vertical 3D slope under two BCs (135° corner): (a) unrestrained; (b) fully restrained.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

Fig. 15. Contours of plastic strain (shear-slip zone) for a vertical 3D slope under fully restrained BC (135° corner).
For personal use only.

90° corner angle under two BCs. Similar plots for vertical Effect of turning arc on a slope
3D slopes with a 135° corner angle under two BCs are The effect of a turning corner on the stability of a 3D
shown in Figs. 14 and 15. slope has been sufficiently analyzed as described above.
Comparing the above two cases (90° and 135° corner), the However, in practical slope engineering, turning corners do
slope-turning corner shows a limited ability to resist potential not exist everywhere, but are replaced by turning arcs in
sliding for a vertical 3D slope under two BCs, even with a some cases. For example, in mountainous areas, some turn-
sharper turning corner (90°corner angle) the constraint effect ing points for cutting across slopes often appear as an arc
is still unappreciated and unexpected. On the contrary, the track in a top view. In the following discussion, a gentle 3D
turning corner of a vertical 3D slope, especially in the 90° slope (1V:2H) with a turning arc of 180° is investigated in
turning corner case, increases the slide risk of the slope. detail. The corresponding top view and side view are shown,
This is exactly why a lower safety factor is defined for a ver- respectively, in Figs. 16 and 6b.
tical 3D slope with a 90° corner angle. The maximum hori- A three-dimensional strength-reduction FEM is used to per-
zontal displacement of a vertical 3D slope is always located form the calculation of slope stability. The safety factors corre-
at the intersection of the turning corner line and the slope sponding to three types of BCs — unrestrained, semi-restrained
crest, which is the most dangerous zone and can be defined and fully restrained — are, respectively, 2.670, 2.685, and
according to the maximum horizontal displacement and 2.690, and the solution of a 2D plane strain analysis is
equivalent plastic strain contours. A potential landslide can 2.555. Comparing the two cases, the solutions obtained
readily occur around this intersection region; it should there- from the 3D analysis are significantly larger than the plane
fore be investigated intensively and effective reinforcement strain solution and also somewhat larger than the values for
should be provided without delay, regardless of the turning a 3D straight slope (turning corner 180°) (Table 5). This is
corner of the vertical 3D slope. because the slope under analysis, shown in Fig. 16, includes

Published by NRC Research Press


Nian et al. 585

Fig. 16. Top view for a gentle 3D slope with turning arc. Fig. 17. Deformed FE mesh for a gentle 3D slope with 180° turning
arc under two BCs: (a) unrestrained; (b) semi-restrained.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

a circular turning arc and certain constraints exist in the 3D tions of the plastic zone (at the turning corner) to restrain
slope, which mean that it no longer satisfies the assump- the sliding of the slope. This indicates that the turning cor-
tions of the plane strain calculation even though the slope ner exerts some constraints or container effects. Therefore,
turning arc is extended over a long perimeter and the slope the FOS for a convex-shaped gentle slope with corner an-
gradient of the points around the turning arc is identical to gle (90°, 135°) is larger than for a 3D straight slope (see
the previous case. Table 5). However, for convex-shaped vertical slopes (e.g.,
Figures 17 and 18 show the deformed mesh and equivalent 90° and 135° corner angle), a major sliding failure surface
plastic strain contours for the gentle 3D slope with turning will cut obliquely across the whole turning corner to the
arc 180° under two BCs. It can be seen from Figs. 17b and base, because high plastic strain zones are concentrated at
For personal use only.

18b that the deformed mesh and shear-slip zone under the the base of the turning corner (see Figs. 13 and 15). The
semi-restrained BC have not been extended to the end of the sliding body will appear as a pyramid-shaped turning cor-
fixed boundary. The reason for this is that the 3D turning arc ner block or one-quarter pyramid. Under this condition,
slope under analysis is somewhat different from the common the turning corner will no longer provide the extra con-
3D straight slopes or 3D slopes with turning corner discussed straint effects, but on the contrary, can readily slide along
above, in which the beeline or fold line occurs along the z- the turning corner block and reduce the stability of the
direction. In general, the present gentle 3D slope with turning convex-shaped vertical slope.
arc 180° under various BCs exhibits certain constraint ef- • Free surface effects that are completely different for convex-
fects, which lead directly to an evident improvement in slope shaped gentle and vertical slope with turning corners —
stability. The steeper the slope inclination, the more marked the
free-surface effects, and the lower the FOS for a 3D con-
Discussion vex slope. In the two cases of convex-shaped gentle and
vertical slopes examined in the paper, the free-surface ef-
A three-dimensional strength-reduction finite element anal- fect of the vertical slope is more evident than that of the
ysis was carried out to obtain the safety factors (FOS) for a gentle one. Moreover, the free-surface effect for a con-
convex slope with a turning corner. An interesting phenom- vex-shaped vertical slope with a 90° corner angle is the
enon regarding the variations in the safety factor as a func- most evident among the three turning corners. Therefore,
tion of the 3D slope surface geometry has been observed. the FOS for a convex-shaped vertical slope with a 90°
The FOS for a convex-shaped gentle slope decreases when corner angle is the minimum (see Table 6).
increasing the corner angle from 90° through 135° to 180° In addition, to further explore the difference between con-
(straight slope). However, the FOS for a convex-shaped verti- vex- and concave-shaped slopes, a concave-shaped vertical
cal slope increases when increasing the corner angle from slope with a 90° corner angle as illustrated in example 2 and
90° through 135° to 180° (straight slope). Fig. 11 was calculated to obtain the FOS and the critical slip
The physical reasons for the observed variations in the surface. The resulting FOS value under various BCs are listed
safety factor as a function of geometry can be interpreted in Table 7, and the deformed mesh and equivalent plastic
through the following two aspects: strain under a fully restrained BC are shown in Fig. 19. Com-
• Different failure modes for convex-shaped gentle and ver- pared with Table 6, the FOS of the concave-shaped vertical
tical slopes with turning corners — For the convex- slope with a 90° corner angle is markedly higher than that of
shaped gentle slope, major sliding failure will occur along a convex-shaped vertical slope with a 90° corner angle. The
the lateral two-slope surface. The plastic shear-strain zone difference between the two cases is close to 15% even though
will cross in the upper middle portion of the turning cor- compared with the 3D straight vertical slope, the maximum
ner (e.g., see Fig. 8b) or the inner portion of the turning difference can also be as high as 8.5% (Tables 6 and 7). The
corner (e.g., see Fig. 10b). Under this condition, when a reason for this is that a completely different sliding failure
lateral slope yields in a deep slide, the other slope will mode and free-surface effect exist in the concave-shaped ver-
produce certain resistance or drag effects at the intersec- tical slope with a 90° corner angle (see Fig. 19b).

Published by NRC Research Press


586 Can. Geotech. J. Vol. 49, 2012

Fig. 18. Contours of plastic strain (shear-slip zone) for a gentle 3D slope with 180° turning arc under two BCs: (a) unrestrained; (b) semi-
restrained.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13
For personal use only.

Summary and conclusions Table 7. FOS for concaved-shaped vertical 3D slopes with 90°
turning corners under various BCs.
Three-dimensional elastic, perfectly plastic FE calculations
with the strength-reduction method have been performed to Safety factor
obtain safety factors and critical slip surfaces for 3D slopes. Turning corner Semi- Fully
In addition, the complex geological conditions (thin weak of slope Unrestrained restrained restrained
layer and groundwater), various boundary conditions, and 90° 1.274 1.275 1.380
complex geometries (turning corners and arcs) are fully con-
sidered. The numerical results obtained have clearly demon- ence in FOS between a gentle slope and a steep one is
strated that the 3D strength-reduction FEM presented in this much greater than 100% (Tables 5 and 6).
paper is reasonable and feasible. In particular, this approach 2. For both gentle and steep 3D slopes with a corner, the
can visually monitor the failure progress of a 3D slope and maximum horizontal displacement is always located at the
clearly outline the critical slip surface over the overall shear intersection of the turning corner line and the slope crest
failure zone in a 3D state, which means that it can be ex- under the fully restrained BC. Here a potential landslide
pected to become more widely popular in geotechnical engi- can readily occur, and effective reinforcement should be
neering practice. From the findings of this study, the pre-emptively carried out. Moreover, the central sliding fail-
following conclusions can be drawn: ure mode shown in this case is entirely different from the
1. For a convex-shaped gentle 3D slope with a corner, signif- lateral sliding failure mode under unrestrained conditions.
icant constraint effects exist at the corner, and therefore 3. The FOS of a concave-shaped vertical slope with a 90° cor-
a considerably higher safety factor is obtained than for a ner angle is markedly higher than that of a convex-
common 3D slope (straight slope). In particular, the shaped vertical slope with a 90° corner angle. The
smaller the corner angle (180°→90°), the larger the maximum difference between the two cases can be as
safety factor, regardless of the boundary conditions. For great as 15% even though compared with the 3D
a steep 3D slope with a corner, however, the constraint straight vertical slope, the maximum difference can also
effects at the corner disappear, and a lower safety factor, be as high as 8.5% (Tables 6 and 7). From the view-
even smaller than that from plane strain analysis, results. point of engineering safety (conservative) and minimiz-
In general, the difference in FOS is less than 10% for a ing computational cost, a concave-shaped vertical slope
gentle 3D slope under a different corner angle or for a with a 90° corner angle can be replaced by a straight
steep one undera different corner angle, while the differ- vertical slope for computation of the safety factor.

Published by NRC Research Press


Nian et al. 587

Fig. 19. Deformation of a concave-shaped vertical 3D slope with 90° turning corner under fully restrained BC. (a) Deformed mesh; (b) con-
tours of equivalent plastic strain.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

4. For a gentle 3D slope with a 180° turning arc, the safety Cai, F., Ugai, K., Wakai, A., and Li, Q. 1998. Effects of horizontal
For personal use only.

factors obtained are significantly greater than the plane drains on slope stability under rainfall by three-dimensional finite
strain solution and also somewhat greater than those ob- element analysis. Computers and Geotechnics, 23(4): 255–275.
tained from a common 3D straight slope. The presented doi:10.1016/S0266-352X(98)00021-4.
3D slope with turning arc shows some constraint effect, Cavounidis, S. 1987. On the ratio of factors of safety in slope stability
which leads directly to an obvious improvement in slope analyses. Géotechnique, 37(2): 207–210. doi:10.1680/geot.1987.
stability. 37.2.207.
Chen, R.H., and Chameau, J.-L. 1985. Discussion: three-dimensional
5. In 3D slope stability analysis, the shape and location of the
limit equilibrium analysis of slopes. Géotechnique, 35(2): 215–
critical slip surface is likely to be different at different
216. doi:10.1680/geot.1985.35.2.215.
points in the z-dimension, especially when the slope geo- Chen, Z., Wang, J., Wang, Y., Yin, J.-H., and Haberfield, C. 2001. A
metry is not regular with a turning corner or turning arc. three-dimensional slope stability analysis method using the upper
bound theorem. Part II: numerical approaches, applications and
Acknowledgements extensions. International Journal of Rock Mechanics and Mining
The research work presented here and the preparation of Sciences, 38(3): 379–397. doi:10.1016/S1365-1609(01)00013-2.
this paper have been financially supported by National Natu- Chen, J., Yin, J.-H., and Lee, C.F. 2003a. Upper bound limit analysis
ral Science Foundation of China (NSFC; Grant of slope stability using rigid finite elements and nonlinear
No. 51179022), the Open Research Fund of the State Key programming. Canadian Geotechnical Journal, 40(4): 742–752.
Laboratory of Geohazard Prevention and Geoenvironmental doi:10.1139/t03-032.
Protection, Chengdu University of Technology (Grant No. Chen, Z., Mi, H., Zhang, F., and Wang, X. 2003b. A simplified
SKLGP2010K005), and the Research Fund of Japan (Scien- method for 3D slope stability analysis. Canadian Geotechnical
tific Research B: 22310113). This financial support is grate- Journal, 40(3): 675–683. doi:10.1139/t03-002.
Cheng, Y.M., and Yip, C. 2007. Three-dimensional asymmetrical
fully acknowledged. The authors also wish to acknowledge
slope stability analysis: extension of Bishop’s, Janbu’s and
the Associate Editor and the two anonymous reviewers for
Morgenstern-Price’s techniques. Journal of Geotechnical and
their helpful comments on the manuscript. Geoenvironmental Engineering, 133(12): 1544–1555. doi:10.
1061/(ASCE)1090-0241(2007)133:12(1544).
References Cheng, Y.M., Liu, H.T., Wei, W.B., and Au, S.K. 2005. Location of
Arellano, D., and Stark, T.D. 2000. Importance of three-dimensional critical three-dimensional nonspherical failure surface by nurbs
slope stability analysis in practice. In Slope Stability 2000. functions and ellipsoid with applications to highway slopes.
Geotechnical Special Publication No. 101. Edited by D.V. Computers and Geotechnics, 32(6): 387–399. doi:10.1016/j.
Griffiths, G.A. Fenton, and T.R. Martin. American Society of compgeo.2005.07.004.
Civil Engineers, New York. pp. 18–32. Cheng, Y.M., Li, L., and Chi, S.C. 2007. Performance studies on six
Bromhead, E.N., and Martin, P.L. 2004. Three-dimensional limit heuristic global optimization methods in the location of critical
equilibrium analysis of the Taren landslide. In Advances in slip surface. Computers and Geotechnics, 34(6): 462–484. doi:10.
Geotechnical Engineering (Skempton Conference). Thomas Tel- 1016/j.compgeo.2007.01.004.
ford, London. Vol. 2, pp. 789–802. Chugh, K.A. 2003. On the boundary conditions in slope stability

Published by NRC Research Press


588 Can. Geotech. J. Vol. 49, 2012

analysis. International Journal for Numerical and Analytical Jiang, J.-C., and Yamagami, T. 2004. Three-dimensional slope
Methods in Geomechanics, 27(11): 905–926. doi:10.1002/nag. stability analysis using an extended Spencer method. Soils and
305. Foundations, 44(4): 127–135. doi:10.3208/sandf.44.4_127.
Dawson, E.M., Roth, W.H., and Drescher, A. 1999. Slope stability Lam, L., and Fredlund, D.G. 1993. A general limit equilibrium model
analysis by strength reduction. Géotechnique, 49(6): 835–840. for three-dimensional slope stability analysis. Canadian Geotech-
doi:10.1680/geot.1999.49.6.835. nical Journal, 30(6): 905–919. doi:10.1139/t93-089.
Duncan, J.M. 1996a. Soil slope stability analysis. In Landslides: Leshchinsky, D., and Baker, R. 1986. Three-dimensional slope
investigation and mitigation. Special Report 247, Chapter 13. stability: end effects. Soil and Foundation, 26(4): 98–110. doi:10.
Edited by A.K. Turner et al. Transportation Research Board, 3208/sandf1972.26.4_98.
Washington, D.C. Manzari, M.T., and Nour, M.A. 2000. Significance of soil dilatancy
Duncan, J.M. 1996b. State of the art: limit equilibrium and finite in slope stability analysis. Journal of Geotechnical and Geoenvir-
element analysis of slopes. Journal of Geotechnical Engineering,
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by University of Melbourne on 05/15/13

onmental Engineering, 126(1): 75–80. doi:10.1061/(ASCE)1090-


122(7): 577–596. doi:10.1061/(ASCE)0733-9410(1996)122:7 0241(2000)126:1(75).
(577). Sainak, A.N. 2004. Application of three-dimensional finite-element
Fredlund, D.G., and Krahn, J. 1977. Comparison of slope stability method in parametric and geometric studies of slope stability. In
methods of analysis. Canadian Geotechnical Journal, 14(3): 429– Advances in Geotechnical Engineering (Skempton Conference).
439. doi:10.1139/t77-045. Thomas Telford, London. Vol. 2, pp. 933–942.
Griffiths, D.V., and Lane, P.A. 1999. Slope stability analysis by finite Stark, T.D., and Eid, H.T. 1998. Performance of three-dimensional
elements. Géotechnique, 49(3): 387–403. doi:10.1680/geot.1999.
slope stability methods in practice. Journal of Geotechnical and
49.3.387.
Geoenvironmental Engineering, 124(11): 1049–1060. doi:10.
Griffiths, D.V., and Marquez, R.M. 2007. Three-dimensional slope
1061/(ASCE)1090-0241(1998)124:11(1049).
stability analysis by elasto-plastic finite elements. Géotechnique,
Ugai, K., and Leshchinsky, D. 1995. Three-dimensional limit
57(6): 537–546. doi:10.1680/geot.2007.57.6.537.
equilibrium and finite-element analysis: a comparison of results.
Hibbitt, Karlsson, and Sorensen Inc. 2005. ABAQUS/Standard,
Soils and Foundations, 35(4): 1–7. doi:10.3208/sandf.35.4_1.
ABAQUS/CAE user’s manual. HKS Inc., Pawtucket, R.I.
Vaughan, P.R. 2008. Discussion on three-dimensional slope stability
Huang, C.C., and Tsai, C.C. 2000. New method for 3D and
asymmetrical slope stability analysis. Journal of Geotechnical and analysis by elasto-plastic finite elements. Géotechnique, 58(8):
683–685. doi:10.1680/geot.2008.D.010.
For personal use only.

Geoenvironmental Engineering, 126(10): 917–927. doi:10.1061/


(ASCE)1090-0241(2000)126:10(917). Wei, W.B., Cheng, Y.M., and Li, L. 2009. Three-dimensional slope
Huang, C.C., Tsai, C.C., and Chen, Y.H. 2002. Generalized method failure analysis by the strength reduction and limit equilibrium
for three-dimensional slope stability analysis. Journal of Geotech- methods. Computers and Geotechnics, 36(1–2): 70–80. doi:10.
nical and Geoenvironmental Engineering, 128(10): 836–848. 1016/j.compgeo.2008.03.003.
doi:10.1061/(ASCE)1090-0241(2002)128:10(836). Zhang, X. 1988. Three-dimensional stability analysis of concave
Hungr, O. 1987. An extension of Bishop’s simplified method of slope slopes in plan view. Journal of Geotechnical Engineering, 114(6):
stability analysis to three dimensions. Géotechnique, 37(1): 113– 658–671. doi:10.1061/(ASCE)0733-9410(1988)114:6(658).
117. doi:10.1680/geot.1987.37.1.113. Zheng, H., Tham, L.G., and Liu, D. 2006. On two definitions of the
Hungr, O., Salgado, F.M., and Byrne, P.M. 1989. Evaluation of a factor of safety commonly used in the finite element slope stability
three-dimensional method of slope-stability analysis. Canadian analysis. Computers and Geotechnics, 33(3): 188–195. doi:10.
Geotechnical Journal, 26(4): 679–686. doi:10.1139/t89-079. 1016/j.compgeo.2006.03.007.
Jeremic, B. 2000. Finite-element methods for three-dimensional slope Zienkiewicz, O.C., Humpheson, C., and Lewis, R.W. 1975.
stability analysis. In Slope stability 2000, GSP 101. Edited by D.V. Associated and non-associated visco-plasticity and plasticity in
Griffiths et al. American Society of Civil Engineers, Reston, Va. soil mechanics. Géotechnique, 25(4): 671–689. doi:10.1680/geot.
pp. 224–238. 1975.25.4.671.

Published by NRC Research Press

Das könnte Ihnen auch gefallen