,, Z J - It Was Observed That For A - The Presented Example Problem Dem

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

GENERALIZED THREE-DIMENSIONAL

SLOPE-STABILITY ANALYSIS
By Dov Leshchinsky, 1 Member, ASCE, and Ching-Chuan Huang 2
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT: A 3-D slope-stability-analysis method, explicitly satisfying all limit-


ing-equilibrium equations, is presented. To make the problem statically determi-
nate, the variational approach, rather than intuition in 3-D, is used. In this ap-
proach, the normal stress over the user's specified slip surface, which satisfies global
limiting equilibrium and produces the minimum factor of safety for the critical
surface, is sought and obtained mathematically. A simple numerical scheme to
attain the safety factor is detailed and all the algorithms needed to assemble a
computer program are given. Upon discretization of the surface by n rectangular
elements, n simultaneous linear equations are solved for n discrete normal stresses.
It is also necessary to solve three nonlinear equations for the factor of safety, Fs,
and the coordinate of the center of rotation [xc, z j . It was observed that for a
selected potential slip surface there may be many possible combinations of (;i +
2) roots, all giving essentially the same Fs. The presented example problem dem-
onstrates the importance of 3-D back-analysis if one is to configure in-situ soil
strength.

INTRODUCTION

Natural and manmade slopes are often inhomogeneous and their ge-
ometry varies in space, even along short distances. Therefore, when slides
occur their pattern assumes three-dimensional geometry. It is common,
however, to assess the margin of safety through a 2-D idealization of the
slope; i.e., an "equivalent" plane-strain problem is postulated and analyzed.
The 2-D simplification is to a large extent intuitive, typically corresponding
to the worst-case scenario. Happily, the end result is usually conservative,
though more expensive, since 3-D effects tend to increase stability. How-
ever, ignoring the 3-D effects in research may lead to unsafe conclusions.
For example, in postfailure studies the exclusion of the third dimension in
back-analysis may imply that the strength of the soil is higher than it actually
was; i.e., based on a 2-D back-analysis of the traced idealized
2-D slip surface, the added stability due to end effects is attributed to the soil
shear strength. A second example has to do with laboratory studies of 2-D
earth structure models at failure. These physical models are contained within
a rigid box of limited breadth. Friction and adhesion are likely to develop
between the soil and the sidewalls of the box, thus enhancing the perfor-
mance of the structure as compared to the ideal 2-D case and possibly leading
to overconfidence. This paper presents a generalized 3-D limit-equilibrium-
analysis method. For a general (but symmetrical) slip surface specified by
the user, the corresponding safety factor can be determined with the si-
multaneous satisfaction of all limiting-equilibrium equations. Subsequently,

1
Assoc. Prof, of Civ. Engrg., Univ. of Delaware, Newark, D E 19716.
2
Assoc. Prof., Dept. of Civ. Engrg., Cheng-Kung Univ., Tainan, Taiwan; for-
merly, Postdoctoral Fellow, Dept. of Civ. Engrg., Univ. of Delaware, Newark, D E .
Note. Discussion open until April 1,1993. To extend the closing date one month,
a written request must be filed with the ASCE Manager of Journals. The manuscript
for this paper was submitted for review and possible publication on May 17, 1991.
This paper is part of the Journal of Geotechnical Engineering, Vol. 118, No. 11,
November, 1992. ©ASCE, ISSN 0733/9410792/0011-1748/$1.00 + $.15 per page.
Paper No. 1909.

1748

J. Geotech. Engrg. 1992.118:1748-1764.


a tool is produced capable of assessing, in a rational manner, the 3-D effects
in problems such as those described in the aforementioned examples.
As is evident from the literature, most of the research effort has concen-
trated on a refinement of 2-D analyses with comparatively little attention
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

given to 3-D aspects of stability analysis. Leshchinsky and Baker (1986)


reviewed some of the 3-D existing methods [e.g., Giger and Krizek (1975,
1976), Lambe and Whitman (1969), Baligh and Azzouz (1975), Azzouz and
Baligh (1976, 1978, 1983), Azzouz et al. (1981), Ugai (1985), Hovland
(1977), Chen and Chameau (1983), Anagnosti (1969), Kopacsy (1957), and
Leshchinsky et al. (1985)]. Some of these referenced methods are limited
a priori to purely cohesive soils. Others fail to degenerate to known solutions
when limiting 2-D cases are considered (Leshchinsky 1990a) or depend on
constants (i.e., lateral earth-pressure coefficients) that are extremely diffi-
cult to estimate in 3-D. Recently, Gens et al. (1988) presented an analysis
for cohesive slopes. Michalowski (1989) introduced, in the strict framework
of limit analysis of plasticity, a rigorous 3-D analysis; however, generali-
zation of this analysis to layered soil has yet to be done. Instructive com-
parative performance of 3-D methods is given by Ugai (1988) and Hungr
et al. (1989).
Leshchinsky and Baker (1986) pointed out that the 3-D variational lim-
iting-equilibrium approach, as presented by Leshchinsky et al. (1985), yielded
failure mechanisms that are too restrictive. They presented a modified 3-D
variational formulation, limited to homogeneous problems, that resulted in
a failure mechanism consisting of a cylindrical body with end caps attached
to it. This surface is a 3-D generalization of the 2-D log spiral function.
However, this failure mechanism cannot be extended to layered soil. Lesh-
chinsky (1990b) has generalized the 2-D variational limit-equilibrium anal-
ysis presented by Baker and Garber (1978) to deal with general slip surfaces.
Leshchinsky and Huang (1992) modified this generalization showing the
variational approach to be equivalent to traditional rigorous limit-equilib-
rium methods, though more efficient. This paper extends the 2-D gener-
alization as presented by Leshchinsky and Huang (1992) to the 3-D stability
analysis presented by Leshchinsky et al. (1985). It provides all the algorithms
necessary for the assembly of a computer program and the required architec-
ture of the computer code, and it demonstrates the solution performance.

FORMULATION

The global geometry of a potentially sliding 3-D mass is illustrated in Fig.


1(a). The z-axis is opposing gravity direction and the region D defines the
projected area of the slip surface on the x-y plane. Fig. 1(b) shows a dif-
ferential area on the slip surface. It is subjected to a normal stress cr and
shear stress T. Since the sliding mass is considered to be in a limit equilibrium
state, T = the "mobilized" Coulomb strength of the soil; i.e., T = (c, +
[<T - w]tan §,)/F where ct and (j>, are the cohesion and internal angle of
friction, respectively, of the soil within which the differential area is located
(soil layer /); F = a reduction factor; and u = the pore-water pressure at
(x, v, z). Note in Fig. 1(b) the angle 6; it is the angle between the x-axis
and the projection of the elemental resultant force of T on the x-y plane.
The direction opposite to T signifies the projected direction of slide move-
ment at (x, y, z). Note that in 2-D problems, where there are no changes
in the y-axis direction, 8 equals 180° and the mass general movement is in
the x-axis direction. Fig. 1(c) defines the scope of the formulated problem.
It is limited to a situation in which both geometry and soil properties are sym-
1749

J. Geotech. Engrg. 1992.118:1748-1764.


Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

FIG. 1. 3-D Sliding Surface: (a) Global Geometry; (b) Differential Area; (c) Sym-
metrical Problem

metrical with respect to a vertical plane. The Cartesisan coordinate system


is selected so that the plane of symmetry coincides with y = 0. Subsequently,
only three out of six equilibrium equations are significant: force equilibrium
in the x and z directions and the moment equilibrium about the y-axis. The
other three equilibrium equations are satisfied by virtue of symmetry. Note
in Fig. 1(c) that only five layers of soil are shown; however, the formulation
is applicable to any number of layers. All three functions z(x, y), cr(x, y),
and 9(x, y) are unknown and are assumed to be continuous.
Considering symmetry and through explicit substitution of variables in
the vectorial equilibrium equations presented by Leshchinsky et al. (1985),
1750

J. Geotech. Engrg. 1992.118:1748-1764.


one can verify, after a laborious process though, that the limiting global
equilibrium equations for the sliding mass contained in layered soil are
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

HX = 2 {[c, + (cr - «H/]# + Fez*} dxdy = 0 (1)


1=1

VZ = 2 {[c, + (a - uMB(zx + ftz„) - F[a - 7(Z - z]} dx dy


I -l J JDl

= 0 (2)

M7 = E {[c, + (o- - «)»|i,]5[z - x(z, + ftz,,)] - F[7*(Z - z)


/ - 1 J JDi

- u{x + zzx)}} dxdy = 0 (3)


where
ft = il(x, v) = tan 9 (4)
<W = tan <)>, (5)

B= (6)
{i + ^Y(z!+wj
and HX, VX, and MY = the equilibrium equations for the horizontal forces
in the x-axis direction, vertical forces in the z-axis direction and moments
about the y-axis, respectively; / = the soil's layer number (/ = 1,2. . . m);
Dt = area of the slip surface, located within layer /, as projected on the x-
y plane [Fig. 1(a)]; z„, zy = partial derivatives of z with respect to x and
y, respectively \zx = dz/dx, zy = dzldy); Z = Z(x, y) = elevation of slope
surface; and 7 = the weighted-average unit weight of the layers of soils in
column (Z-z).
From (l)-(3) it is clear that F = a function of z(x, y), u(x, y), and %{x,
y). Hence, F = a functional. The limiting-equilibrium analysis seeks the
minimum value of F, that is, the factor of safety Fs. To achieve this, the
functions z(x, y), a(x, y), and Q(x, y) that realize Fs = min(.F) must be
known. Finding these critical functions can be done through the variational
calculus by using any one of the three equilibrium equations to minimize
F. However, the other two equilibrium equations must also be satisfied for
F = Fs and its associated z(x, y),a(x,y), and Q(x, y). Therefore, the other
two equations are considered constraints. Using HX to define F, and VZ
and MY as constraints, and expanding to layered soil the approach presented
by Leshchinsky et al. (1985) [who extended the 2-D procedure introduced
by Baker and Garber (1978)], one can show the problem to be equivalent
to the minimization of an auxiliary functional, G, defined as

G = X I g dx dy (7)
1=1 J JDi

where
g = [c, + (ex - u)ty,]B{l + \(zx + ftzr) + |x[z - x(zx + ilzy)]}
+ F > [ z v - \ + (JL(X + zzx)} + ^(Z - z)(\ - iix)} (8)

1751

J. Geotech. Engrg. 1992.118:1748-1764.


and X, JJL = Lagrange multipliers. Minimization of G yields the sought
functions z(x, y), a(x, y), and 8(x, y), containing three unknown parameters
X, (x, and Fs. Back-substitution of these functions into the three equilibrium
equations [(l)-(3)] enables one to determine Fs, as well as X and (JL.
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

It can be verified [e.g., Leshchinsky (1990b)] that if the two Lagrange


multipliers are replaced by xc = X/(JL and zc = — l/u>, (8) degenerates to
the integrand in the moment equilibrium equation [(3)] as written about a
translated y-axis at (xc, zc). This implies that the minimization process is
equivalent to minimizing F using only the moment equilibrium equation
written about the y-axis at the unknowns (xc, zc). Hence, Lagrange multi-
pliers can be viewed as a coordinate defining a fictitious axis of rotation for
the sliding mass (Leshchinsky and Huang 1992).
Based on the calculus of variations, the critical functions z(x, y), tr(x, y),
and 0(x, y) (or ft = ft[x, y] = tan 9) that minimize the functional G, and
therefore produce the factor of safety Fs, should satisfy the Euler's equations
[e.g., Elsgolc (1962)]:

*L-JL(°L)-±(1L)=O (9)
K
dft dx \dilj dy \my) '

--•f(*L)--f(*L)=o dvo)
do- dx \doJ dy \d(jyJ '
^ d / 3 g \ d / d S \ = Q
v
dz dx \dzj dy \dzyJ '
where ftx = 3ft/to, Q,y = 3ft/5y, ax = da/dx and <ry = d&ldy. Hence, the
application of Euler's equations should enable one to determine fl(x, y)
and cr(;t, y).
Combining (9) with (8), substituting xc = X/u. and zc = -1/u., and
rearranging the terms yield
z
ft = tan 6 = y(Xc - *) - zxzy(z - ze)
zx(xc - x) + (z - zc) + Zy\z - zc)
Combining (10) with (8) will yield the differential equation for the slip
surface z(x, y). Leshchinsky et al. (1985) and Leshchinsky and Baker (1986)
have successfully solved this equation for homogeneous soil. The surface
then is a 3-D generalization of the 2-D log spiral. For this surface, Fs can
be determined without specifying <r(x, y) since the moment-equilibrium
equation written about (xc, zc) is independent of the normal stress (Lesh-
chinsky and Baker 1986). Leshchinsky et al. (1985) showed this surface to
yield a solution equivalent to an upper bound in the context of limit analysis
of plasticity. However, this mathematically derived surface is extremely
restrictive, i.e., it cannot be extended to realistic situations where layered
soil is likely to dictate its trace. Subsequently, similar to the 2-D generali-
zation suggested by Leshchinsky (1990b), z(x, y) in this paper is left to be
specified by the user. This approach is common to all generalized methods
of 2-D slope-stability analysis where one of several available optimization
techniques is used to locate the critical surface. Note that by specifying z(x,
y), 0 in (12) contains only two unknown parameters: xc and zc. Back sub-
stitution of 6(x, y) and z(x, y) into the equilibrium equations [(l)-(3)] gives
three equations with three unknown parameters (xc, zc, and Fs) and one
1752

J. Geotech. Engrg. 1992.118:1748-1764.


unknown functionCT(X,y). If the stress function is obtained, then the three
unknown parameters can be determined. Variation of z(x, y) by the user
will eventually produce the critical slip surface rendering minimum Fs.
Application of (11) to (8) results in the partial differential equation for
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

the critical a(x, y) containing over 100 terms. Although this expression was
developed with the aid of an advanced computer code capable of dealing
with symbolic algebraic derivations, the writers felt uncomfortable in uti-
lizing it. Instead, the terms dg/dz, dg/dzx and dg/dzy were derived manually
in a straightforward manner and integrated into a numerical procedure that
satisfies (11) and thus producing discrete values of cr(x, y), as shown in the
next section. For completeness, these terms are given next:

fz = [c, + (CT - u)^,]B + FS<JZX + Fsy(x - xc) (13)

dg
— = [c, + (CT - u)tyt]B(xc - x) + Fscr(z - zc) + [c, + (CT - u)\\>,]

•[(z - zc) - (x - xc)(zx + flz,)] — (14)

de
-£• = [c, + (CT - u)ty,]B(xc - x)Q + [c, + (CT - i*H,][(z - zc)

- (x - xc)(zx + Ozy)] — (15)

where:
35 = zx{\ + ft2) + (zx + £lzy)[zy(zxn - zv) - 1]
dzx {z2x + z2y + 1)1/2[1 + il2 + (zx + ilzy)2]3/2 {
'
and
dB _ zy{\ + a2) + (zx + Q,zy)[zx(zy - zJX) - O,]
(17)
dzy (z2 + z 2 + i) 1/2 [i + a 2 + (zx + cizy)2Y
Substitution of (13)-(17) into (11) and carrying out the total derivatives
with respect to x and y will result in the differential equation for u(x, y).
One can visualize the complexity of this equation. However, differentiation
with respect to x and y in the numerical scheme shown in the next section
is rather simple.
It should be emphasized thatCT(X,y) is basically a term of the integrands
in the equilibrium equations [(l)-(3)] and only its integrated value must
satisfy these global equations. That is, in the framework of limit-equilibrium
formulation, <j(x, y) does not have to satisfy local stress-equilibrium re-
quirements (e.g., dajdx + dixy/dy + dx^/dz + X = 0, etc., are not satisfied
along the surface). Hence, although the resultant vector associated with the
variationally obtainedCT(X,y) should satisfy global equilibrium and produce
a unique Fs, there are many possible distributions of cr(x, y) capable of
generating the same resultant (Leschinsky and Huang 1992). Consequently,
CT(X, y) enables one to introduce Coulomb's failure criterion in the for-
mulation and thus to attain the fundamental objective: m i n ^ ) while sat-
isfying global limiting equilibrium. As a result, judgment about internal
forces (i.e., location, inclination, and magnitude) that correspond to the
1753

J. Geotech. Engrg. 1992.118:1748-1764.


variationally obtained <r (x, y) is not necessary in order to attain the objective
of the analysis. This is unlike other rigorous methods where such judgment
is recommended as part of the reasonableness assessment of the analysis
assumptions. However, to assure satisfaction of Coulomb's failure criterion,
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

T(JC, y) must be positive, implying that (cr - u) > -(c/tan 4>), everywhere
over the slip surface. Alternatively, if a truncated failure envelope is used,
one can impose (CT - u) > k where k = the tensile strength cutoff discussed
in the next section.

NUMERICAL PROCEDURE

The 3-D slope-stability analysis presented is limited to symmetrical prob-


lems and therefore, only one-half of the sliding mass needs to be considered
in the equilibrium equations. Fig. 2 shows a plan view of half the slip surface
specified by the user. This surface is discretized by n rectangular elements
with sides parallel to the x- and y-axes. Fig. 3 illustrates the computational
process. Upon discretization of z(x, y), the partial derivatives zx and zy are
calculated at the center of each element (i, j) following a simple finite-
difference scheme. Next, initial values for xc, zc, and Fs are assumed. Now,
Cl(i, j) (or 9[i,;']) and B{i, j) are calculated using first (12) and then (6).
Eqs. (13)-(17) are utilized to compute the partial derivatives dg/dz, dg/dzx
and dg/dZy at all centers [i, j). Note that these terms include the unknowns
CT((, ;'). Since u(i, j) must satisfy (11) (Euler's equation) everywhere, this
equation can be invoked n times and hence, provide n linear equations for
the n unknowns a(i, j). However, the explicit expressions for the total
derivatives [i.e., the second and third terms in (11): d( )ldx and d{ )/dy] are
not given. It is suggested to use the following approximated expressions
(available in numerical handbooks) for these derivatives:

+ {al ~a\
d__
a\
dzx
-w
">i£ -i./
(18)
dx dzx axa2{ax + a2)

Projection of discretized
/ slip surface on x-y plane.
I.I 1,2
2,1 2,2 1,3

3,1 3,2 2,3


3,3
—&£-
5,3
7,1 7,2

Vy

FIG. 2. Plan View of Discretized Slip Surface


1754

J. Geotech. Engrg. 1992.118:1748-1764.


* ^ D I S C R E T I Z E A SLIP SURFACE!

I CALCULATE AT (ij) : zx = dz/dx, zv = dz/dy]


Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

*
-**H SPECIFY a0

1
| ASSUME INITIAL VALUES: F„ xc,~z~^m
1 j

t. , f
j USE EQS. 12 & 6 TO CALCULATE fi(i, j ) & B{i,j)\ |

"IT
COMBINE EQS. 13-23 WITH EQ. 11. SOLVE THE
RESULTED n SIMULTANEOUS LINEAR EQS. FOR a(i,j)

CONDUCT NUMERICAL INTEGRATION OF EQS. 1-3

Fj, xc, zc [and cr(i, j j j O B T A I N E D ]

"YES-

-YES

(^This is done automatically by a routine that solves three simultaneous nonlinear


equations.
^a(i,j) — <r(i:j) — v-{hj) and k — user's specified tensile stress cutoff.
Note: Coulomb's criterion set the lowest limit as k = ~~(c/ tan (f>)t-

FIG. 3. Computational Scheme

*>l\TT {bl-b\)[^\ -b\\^-


~±(s£ dz
y/i.i+i dzyi ij az.yi ij-i
.(19)
Ay \tey)\u, blb2(b1 + b2)

where
1755

J. Geotech. Engrg. 1992.118:1748-1764.


fli = xL] - x,-Lj (20)
a2 = xi+uj - x,v (21)
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

bi = y, v - y,.i-i (22)
b2 = y , v + 1 - y.-.j (23)
Combining (13)—(15) and (18)-(19) with (11) gives n linear equations with
n unknown u(i, /'). Solving these simultaneous linear equations produces
cr(/,;). Now, the integration of the three equilibrium equations [(l)-(3)]
can be carried out numerically in a straightforward manner [similar to Lesh-
chinsky and Huang (1992)]. If these equations are satisfied then the initially
assumed values xc, zc, and Fs are correct; if not, however, new values are
assumed (using a routine that solves three simultaneous nonlinear equations)
and the process is repeated until equilibrium is satisfied. Subsequent to the
aforementioned description, the problem is reduced to (3 + n) unknowns
[i.e., xc, zc, Fs, and n times a(i, /)] that are obtained through solving n
simultaneous linear equations and three simultaneous nonlinear equations.
Note in Fig. 3 the term cr0; the location of this normal stress is at the
crest and it is denoted by point 0 on Fig. 2. This stress is introduced into
the numerical scheme through (11) when calculating the terms at point (1,
1). Introduction of rr0 enables the user to somewhat control a(r, j) distri-
bution and thus to produce compliance with Coulomb's failure criterion, if
necessary (i.e., see last section: [CT - u\uj a k). Since tensile stress tends
to develop at the crest, the authors have found that specifying the normal
stress at point 0 is effective in producing this compliance. The user should
specify various values of cr0 until minimum Fs and admissible stresses over
the entire slip surface are obtained [e.g., Leshchinsky and Huang (1992)].
If for all specified <r0 the resulted stress at some point is (<r - u)uj < k, a
tension crack should then be introduced. This situation implies that the
selected slip surface is statically unfeasible.
The computation process should be repeated for all feasible slip surfaces
until the minimal factor of safety is obtained. The results then (including
the slip surface) are considered critical.

RESULTS
To demonstrate the nature of the 3-D solution, a few results are presented.
Additional results are presented in Chowdhury (1992). Rather than intui-
tively specifying an arbitrary 3-D slip surface for this demonstration, it seems
useful to select one that is also a result of the nongeneralized variational
analysis. That is, to choose an extended log-spiral mechanism for which Fs
can be determined through a different procedure as well [see paragraph
following (12)]. Selection of such a surface should provide the reader with
a visual sense of critical variational results as related to homogeneous prob-
lems.
Figs. 4(a)-4(c) show half of the extended log-spiral surfaces obtained for
conical heaps inclined at 30° (Baker and Leshchinsky 1987). Fig. 4(a) illus-
trates the critical slip surface for a c-4> soil while Figs. 4(b)-4(c) represent
the surfaces for cohesive soils. Fig. 4(b) indicates a deep-seated failure for
homogeneous heap and foundation soils. Fig. 4(c) demonstrates the effects
of impenetrable foundation on the critical slip surface and the required
stability number N. Note that 3-D surfaces tend to deepen as the ratio (Nl
tan 4>) increases, similar to the trend resulting from 2-D analysis. Fig. 5(a)
1756

J. Geotech. Engrg. 1992.118:1748-1764.


Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

FIG. 4. Extended Log Spiral Critical Slip Surface for 30° Conical Heaps: (a) c-<(>
Soil; (b) Cohesive Soil and Deep-Seated Failure; (c) Cohesive Soil over Impene-
trable Foundation

shows the critical surface predicted for a vertical corner cut comprised of
soil possessing high-friction angle (Leshchinsky and Baker 1986). Fig. 5(b)
shows a plan view of the surface in Fig. 5(a), as well as the direction of the
shear resistance force over the sliding surface. This direction results from
the solution of (12) and, physically, it is in the opposite direction to the
slide movement. Only when t)> = 0° all elemental shear forces over the
extended log spiral surface are parallel (i.e., 0 = 180°).
1757

J. Geotech. Engrg. 1992.118:1748-1764.


Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

FIG. 5. Vertical Corner Cut: (a) Critical Slip Surface; (b) Plan View of Direction of
Shear Resistance over Slip Surface

A problem against which the numerical procedure was tested is shown


in Fig. 6. The critical results for the extended log spiral were determined
(Leshchinsky and Mullett 1988) using the alternative formulation (i.e., the
nongeneralized one) while restricting the total failure length not to exceed
2H. Subsequently, xc, zc, and Fs resulting from the present numerical pro-
cedure could be verified for accuracy. Note that the resulted surface is
comprised of a central cylinder with smoothly attached end caps. The slip
surface was discretized as shown in Fig. 7 to apply the numerical procedure.
Fig. 7(a) shows a plan view of the soil "columns" and Fig. 7(b) illustrates
cross sections for the ylH values defining the sides of the columns in the y
direction. Note that 20 elements were chosen after the writers realized no
significant difference in results when 61 columns were used (10 of which
were along the symmetry plane). Fig. 8 illustrates the normal stress distri-
bution over the slip surface resulted from the numerical analysis, a function
that remains unknown in the alternative formulation. Although this stress
1758

J. Geotech. Engrg. 1992.118:1748-1764.


Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

FIG. 6. Critical Slip Surface for Example Problem

appears "reasonable" along cross sections of the x-z plane [Fig. 8(a)], it is
perhaps unreasonable (i.e., wavey) along sections on the y-z plane [e.g.,
Fig. 8(£>)]. However, the writers obtained several solutions yielding different
combinations of xc, zc, and cr(7,/) with essentially the same Fs. Subsequently,
it is possible that one of these solutions will improve the appearance of cr {x,
v), though without affecting the analysis objective; i.e., m i n ^ ) . Referring
to the discussion in the second paragraph following (17), one realizes that
the actual value of <T(X, y) does not warrant examination except if it violates
the failure criterion.
The factor of safety obtained for the mechanism and problem shown in
Figs. 6 and 7 is 1.31 and it corresponds to v0/yH = -0.13. Since $u = 0,
most 2-D slope-stability-analysis methods will yield identical predictions
when circular slip surfaces are considered. Hence, one can verify that Fs =
1.00 for the equivalent 2-D problem. The traces of the critical 2-D and 3-
D (on the symmetry plane) slip surfaces are shown in Fig. 9. Conducting
2-D stability analysis [e.g., Leshchinsky and Huang (1992)] on the critical
trace of the 3-D problem shown in Fig. 9 yielded a safety factor of 1.01,
nearly the critical 2-D value of 1.00. Consequently, the writers feel this
problem to be instructive, demonstrating the importance of 3-D back-anal-
ysis in post failure investigation. For example, consider a hypothetical case
in which the exact failure pattern shown in Fig. 6 has happened. Selecting
the deepest traced surface, which is also cylindrical to an extent of 0.4H,
for 2-D analysis will yield a factor of safety of about one (i.e., 1.01), thus
explaining the failure and justifying the selection of c„. However, conducting
a 3-D analysis over the entire traced surface will yield a factor of safety of
1.31, implying that failure occurred because N = 0.261/1.31 = 0.2; i.e.,
the actual cu is about 25% lower than initially estimated. Since back-analyses
of failure are often used to assess the adequacy of laboratory and field
1759

*
J. Geotech. Engrg. 1992.118:1748-1764.
J^z/H-
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

13
-0.2-•
Lf-20

0.2 0.4 o.6 0.8 i.o via

x/H "0- zjK = 0.0


<<•)
z/H ,
u

1 I SSW-WBW
#/ / /
«<&/
/ // // ////
1

0.4'
7 foV
v #/
V/
02-

;;
<—4 j i — —^
x/H -0.2 -0.4 -0.6
C6;

FIG. 7. Discretized Slip Surface for Example Problem: (a) Plan View of Soil "Col-
umns"; (h) Cross Sections

procedures to determine in situ shear strength (e.g., UUversus Vane shear


tests and peak versus residual strength parameters), this example shows that
ignoring 3-D effect may lead to procedures that overestimate the soil's in
situ shear strength. With the combination of both overestimated shear strength
and prevailing 2-D conditions (e.g., failure longer than 2H can develop),
safety is jeopardized even though 2-D analysis is likely to be conducted.
Finally, to demonstrate the capability of the generalized approach to deal
with layered c-§ soil and effective stress analysis, the profile shown in Fig.
10 was utilized. Here, the same slip-surface discretizations as in Fig. 7(a)
was used. For uJyH = -0.10 and zero pore-water pressure, the resulting
factor of safety was 1.14. Using pore-water pressure defined by the param-
eter ru - 0.1, the factor of safety dropped to 1.04. Next, the soil layers in
Fig. 10 were replaced with soils having N = 0.261 and c> | = 0° for the upper
layer, and N = 0.261 and <J> = 30° for the lower layer. For val~iH = -0.13
and ru = 0.0, 0.1 and 0.2, the resulting factors of safety were 1.82, 1.75,
and 1.61, respectively. It should be emphasized that the 3-D slip surface
used is not necessarily the critical one for the two-layered soil problem.
1760

J. Geotech. Engrg. 1992.118:1748-1764.


Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

FIG. 8. Resulted Normal Stress Distribution: (a) Along x-z Cross Sections; (b)
Along One y-z Section

CONCLUSIONS

A slope-stability method, capable of dealing with a general but sym-


metrical 3-D slip surface, is presented. The method is based on the varia-
tional limit-equilibrium approach. It is an extension of Leshchinsky and
Huang's (1992) 2-D analysis and is rigorous in the sense that all global
limiting-equilibrium equations are explicitly satisfied. This is obtained through
a mathematical process in which the normal stress over the specified slip
surface is part of the solution. Hence, the need for statical assumptions in
3-D, where intuition has not been gained yet, is eliminated.
The proposed numerical procedure has been tested against several prob-
lems, one of which is fully reported. It involves solving n simultaneous linear
equations and three nonlinear ones. It was observed that there are several
possible combinations of roots for these equations, all basically giving the
same factor of safety. This multiple roots phenomenon should not pose a
problem since the analysis' objective is to find the safety factor for a given
1761

J. Geotech. Engrg. 1992.118:1748-1764.


CRITICAL 3 - D - y ^ p R i T I C A I. 2-1)
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

F,(2- D) = 1.00

4> = o "

N = — = 0.2610
•ytl

'Total length of failure = 2 . 0 / /

FIG. 9. Traces of Critical Slip Surfaces

kz/H

1. ru = 0 1-00 " ^ ^ 5 ? 5 5 5 ^ ^ ^ ^ ^

2. r u = 0.1
iV = — = 0.1 4> = 30°
7±/
0.49

iV = — - = 0.208 <f> = 15°


z/#
?mmm
FIG. 10. Layered Soil for Extended Example Problem

surface. However, the convergence criteria specified in the computer routine


that solves the three nonlinear equations may have significant effects on the
roots, including the factor of safety. This numerical aspect warrants further
studies that are beyond the scope of the presented paper. Modification to
include external loads can be done by updating only the three equilibrium
equations; i.e., all other equations are not affected directly by such loads.
The example problem presented demonstrates the significance of 3-D
back-analysis. If end-effects are ignored, the in situ soil strength calculated
through local analysis can be overestimated. This may lead to unsafe con-
clusions regarding the adequacy of laboratory and field test results to be
used in stability analyses.
1762

J. Geotech. Engrg. 1992.118:1748-1764.


ACKNOWLEDGMENT

This project is supported by the National Science Foundation under Grant


No. CES-8722818. This support is gratefully acknowledged.
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

APPENDIX. REFERENCES
Anagnosti,P. (1969). "Three-dimensional stability of fill dams." Proc. 7th Intl. Conf.
on Soil Mech. and Foundation Engrg., International Society of Soil Mechanics and
Foundation Engineering, 2, 275-280.
Azzouz, A. S., and Baligh, M. M. (1976). "Design charts for three-dimensional
stability of cohesive slopes subjected to surcharge loads." Publication No. PB-
275779, National Technical Information Service, U.S. Dept. of Commerce, Spring-
field, Va.
Azzouz, A. S., and Baligh, M. M. (1978). "Three-dimensional stability of slopes."
Publication No. PB-285740, National Technical Information Service, U.S. Dept.
of Commerce.
Azzouz, A. S., and Baligh, M. M. (1983). "Loaded areas on cohesive slopes." /.
Geotech. Engrg. Div., ASCE, 109(5), 726-729.
Azzouz, A. S., Baligh, M. M., and Ladd, C. C. (1981). "Three-dimensional stability
analysis of four embankment failures." Proc. 10th Intl. Conf. on Soil Mech. and
Foundation Engrg., 3, 343-346.
Baker, R., and Garber, M. (1978). "Theoretical analysis of the stability of slopes."
Geotechnique, 28(4), 295-411.
Baker, R., and Leshchinsky, D. (1987). "Stability analysis of conical heaps." Soils
Found., 27(4), 99-110.
Baligh, M. M., and Azzouz, A. S. (1975). "End effects on stability of cohesive
slopes."/. Geotechnical Engrg. Div., ASCE, 101(11), 1105-1117.
Chen, R. H., and Chameau, J. L. (1983). "Three-dimensional limit equilibrium
analysis of slopes." Geotechnique, 32(1), 31-40.
Chowdhurry, S. (1992). "Generalized slope stability analysis," MCE thesis, Uni-
versity of Delaware, Newark, Del.
Elsgolc, L. E. (1962). Calculus of variations. Pergamon Press, Ltd., London, Eng-
land.
Gens, A., Hutchinson, J. N., and Cavounidis, S. (1988). "Three-dimensional analysis
of slides in cohesive soils." Geotechnique, 38(1), 1-23.
Giger, N. W., andKrizek, R. J. (1975). "Stability analysis of vertical cut with variable
corner angle." Soils Found., 15(2), 63-71.
Giger, M. W., and Krizek, R. J. (1976). "Stability of vertical corner cut with con-
centrated surcharge load." J. Geotech. Engrg. Div., ASCE, 102(1), 32-40.
Hovland, H. J.(1977). "Three-dimensional slope stability analysis method." /. Geo-
technical Engrg. Div., ASCE, 103(9), 971-986.
Hungr, O., Salgado, F. M., and Byrne, P. M. (1989). "Evaluation of three-dimen-
sional method of slope stability analysis." Can. Geotech. J., 26(4), 679-686.
Kopacsy, J. (1957). "Three-dimensional stress distribution and slip surface in earth
work at rupture." Proc. 4th Intl. Conf. on Soil. Mech. and Foundation Engrg.,
International Society of Soil Mechanics and Foundation Engineering, 1, 339-342.
Lambe, T. W., and Whitman, R. V. (1969). Soil mechanics. John Wiley Book Co.,
New York, N.Y., 370.
Leshchinsky, D. (1990a). Discussion of: "Three-dimensional stability analysis of
concave slopes in plan view," by Zhang Xing, /. Geotech. Engrg., ASCE, 116(2),
342-345.
Leshchinsky, D. (1990b). "Slope stability analysis: Generalized approach," /. Geo-
tech. Engrg., ASCE, 116(5), 851-867.
Leshchinsky, D., and Baker, R. (1986). "Three-dimensional slope stability: End
effects," Soils Found., 26(4), 98-110.
Leshchinsky, D., Baker, R., and Silver, M. L. (1985). "Three-dimensional analysis
of slope stability," Intl. I. Nwner. Anal. Methods Geomech., 9(2), 199-223.
Leshchinsky, D., and Huang, C. C. (1992). "Generalized slope stability analysis:
interpretation, modification and comparison."/. Geotech. Engrg., ASCE, 118(10).
1763

J. Geotech. Engrg. 1992.118:1748-1764.


Leshchinsky, D., and Mullett, T. L. (1988). "Design charts for vertical cuts," / .
Geotech. Engrg., ASCE, 114(3), 337-344.
Michalowski, R. L. (1989). "Three-dimensional analysis of locally loaded slopes."
Geotechnique, 39(1), 27-38.
Downloaded from ascelibrary.org by UNIVERSIDADE FEDERAL DO RIO on 04/22/14. Copyright ASCE. For personal use only; all rights reserved.

Ugai, K. (1985). "Three-dimensional stability analysis of vertical cohesive slopes."


Soils Founds., 25(3), 41-48.
Ugai, K. (1988). "Three-dimensional slope stability analysis by slice methods." Proc.
6th Intl. Conf. on Numerical Methods in Geomech., G. Swoboda, ed., International
Committee for Numerical Methods in Geomechanics, A. A. Balkema, 1369-1374.

1764

J. Geotech. Engrg. 1992.118:1748-1764.

You might also like