You are on page 1of 100

KEYNOTE LECTURE GEOTECHNICAL ENGINEERING OF

THE STABILITY OF NATURAL SLOPES, AND CUTS AND FILLS


IN SOIL
Robin Fell1, Oldrich Hungr2, Serge Leroueil3 and Wynfrith Riemer4

ABSTRACT
The paper presents a critical review of the state of the art of geotechnical engineering of natural slopes, cuts and fills
in soil. Topics which are covered include the pre and post failure mechanics at the micro and macro scale, including
discussion of contractant and other strain weakening soils, creep, progressive and retrogressive failure and fissured
clays. Geotechnical investigation requirements and methods for analysis of stability and deformations, and for analysis
of post failure velocity and travel distance are reviewed.
It is concluded that many slope instability hazards may be managed by traditional factor of safety methods, but that
it is important that the post failure behaviour be considered. Observational, and risk assessment methods may be more
appropriate than the traditional methods in many cases.

1.

INTRODUCTION

1.1

General Objectives of the Paper


This paper sets out to present a critical review of the state of the art of geotechnical engineering of natural slopes,
cuts and fills in soil. This includes site characterization, including the geology and hydrogeology, the establishment of
the material properties and pore pressures, the mechanics of sliding at the micro and macro scale, analysis of stability,
deformations pre and post failure, and management of the slope hazard. The emphasis is on practical issues, and the
integration of engineering geology, hydrogeology, rock and soil mechanics. We will be seeking to demonstrate that it is
now practical to quantify the potential post failure deformations sufficiently well that their assessment should be part of
any geotechnical assessment of a slope. In this way, the engineering of a slope can be linked to the consequences of
failure, allowing better management of the risks. With this in mind, we have emphasized the need to identify and
properly analyze those situations which have the potential to result in large and possibly rapid post failure movements,
which may result in damage to property and loss of life. We have also emphasized those situations which are not well
modelled by conventional effective stress analyses.
1.2
Terminology and Landslide Classification
1.2.1
General
One of the purposes of this conference is to promote better interaction and understanding between various branches
of engineering geoscience. The need for such an effort becomes very apparent to any student of landslide classification
and terminology, since different groups have formulated their own separate vocabulary.
Engineering geologists, being concerned primarily with landslide identification and hazard assessment, have
developed a series of descriptive classification systems. Baltzer (1875) was the first to set out falls, slides and flows as
the principal movement mechanisms. Notable advances in landslide classification were made by Howe (1909), Heim
(1932) and Sharpe (1938). The two classification systems most widely used in the English speaking world today,
Varnes (1978) and Hutchinson (1988) are a direct outgrowth of this long development process. Cruden and Varnes
(1996) revised Varnes system, introducing a systematic taxonomic approach (see below).
In the meantime geotechnical engineers, who are usually concerned with stability assessment and slope stabilization,
seem to have concentrated less on descriptive factors and more on the characteristics and behaviour of materials. Their
attitude was typified by Terzaghi (1950), who turned down an opportunity to formulate yet another classification
system in favour of discussing the various factors predisposing a slope to failure. Similar summaries of relevant
1

Robin Fell, Professor, School of Civil and Environmental Engineering, University of New South Wales, Sydney,
Australia 2052.
2
Oldrich Hungr, Associate Professor, Department of Earth and Social Sciences, University of British Columbia,
Vancouver, BC, Canada V6T 1Z4.
3
Serge Leroueil, Professor, Department de Gnie Civil, Facult des Sciences et de Gnie, University Laval, Quebec,
Canada G1K 7P4.
4
Wynfrith Riemer, Geological Consultant, Risch Haff, L-85 29 Ehner, Redingen, G.D. de Luxumbourg.

factors, especially material properties, have been prepared by Skempton and Hutchinson (1969), Hutchinson (1988) and
others. A systematic problem oriented classification of soils was proposed by Morgenstern (1992).
It is natural that engineers and geologists should have different perspectives on terminology and classification.
However, usage by the two groups has not always been complementary, but separate and parallel. Many concepts held
important by one group are neglected by the other.
A good example of this is the approach to the classification of soil material involved in landslides. Some
geological engineering classifications divide soils into two groups, earth - in which 80 percent or more of the particles
are smaller than gravel (2mm diameter) and debris with more than 20% of gravel and coarser (Varnes, 1978, Cruden
and Varnes, 1996). For geotechnical engineers, the primary division is between cohesive and cohesionless soils, which
have fundamentally different mechanical and drainage characteristics (Morgenstern, 1992). There is almost no
relationship between these two fundamental groupings. Poorly sorted debris with as much as 65% coarse grain content
by volume can be cohesive, when supported by a clayey matrix (Rodine and Johnson, 1976). Laboratory tests by Holtz
and Ellis (1968) showed that adding up to 35% by weight of gravel to a well-graded clayey soil resulted in no
measurable change of the effective friction angle and cohesion. Therefore, the limiting 20% content of coarse clasts,
used in the engineering geological classifications, has little significance with respect to the mechanical behaviour of the
soil. Similar discrepancies exist in the understanding of other factors such as failure mechanisms, saturation and porepressure changes by the two groups.
The following discussion of landslide classification and terminology attempts a balanced view, based on both
geological and engineering outlook.
1.2.2
Terminology
A standard set of terms for description of landslides has been suggested by the Commission on Landslides and
Other Mass Movements of the International Association for Engineering Geology and the Environment (IAEG 1990)
and the UNESCO Working Party on the World Landslide Inventory (WP/WL1). This includes a general definition of a
landslide, which contrasts with an earlier perception that the word landslide specifically implies sliding movement
(Cruden, 1991):
Landslide - the movement of a mass of rock, debris, or earth (soil) down a slope (under the influence of gravity).
The standards also define a large number of useful terms describing the dimensions and morphology of landslides.
Many of these are summarized in Figure 1. These terms appear to have been equally well accepted by both geologists
and engineers. The rupture surface is often called sliding surface in engineering literature, reflecting a
preoccupation with sliding failure mechanisms.
Some of the morphological terms are difficult to apply to landslides with long runout such as debris flows or debris
avalanches, where the displaced mass completely vacates the space above the rupture surface and moves down the
slope. The space between the rupture surface and the original ground could then be called source. This is a useful
addition to the zone of depletion for all landslides. The strip of land traversed by the moving mass down-slope from
the source may be referred to as path or trail, while the accumulation zone in this case coincides with the
deposit as shown in Figure 2 (King, 1999). The word runout, can be used to describe the distance between the toe
of the source and the toe of the deposit. The term travel distance is defined to describe the distance from the crown
of the rupture surface to the toe of the deposit (L in Figure 3). The travel distance angle, a, is defined as shown in
Figure 3.

Figure 1 : Terms describing the morphology of a landslide (Cruden and Varnes, 1996)

Terminology of landslide velocity was formulated by Varnes (1978) and revised by Cruden and Varnes (1996),
resulting in a seven class scale shown in Table 1. Hungr (1981(a)) suggested that velocity is the most important
measure of hazard intensity (cf. Morgenstern (1985)). Thus, the revised velocity classes can be correlated with certain
patterns of human response to landslide hazard (Column 5 of Table 1).

SOURCE

PATH
DEPOSIT

Figure 2 : Components of a long-runout landslide (concept after King, 1999).

Figure 3 : Definition of travel distance (L) and travel distance angle a


R ock avalanche
D ebris avalanche
M ud flow
D ebris flow
E arth flow
C lay flow -s lide
S and,silt, debris flow -slide

1 E -6

0 .0 0 0 0 1 0 .0 0 0 1

0 .0 0 1

0 .0 1

0 .1

10

10 0

V elo city (m /sec)

Figure 4 : Collection of flow-like landslide velocity values (Hungr et al,2000)


It is unfortunate that the terms of the velocity scale in Table 1 have been assigned to velocities that are apparently
too slow. Figure 4 shows a random collection of landslide velocities, grouped by classes. It appears that only a fairly
small fraction of landslide types is limited to velocities less than 5m/sec. Therefore, most flow-like landslide types
carry the somewhat clumsy label of extremely rapid. Despite this, many references to high velocity landslides refer
to them as rapid (e.g.rapid earth flows of Varnes, 1978) even though technically this would limit their velocity to
only a few metres per minute. It probably would not be productive to try to change the terminology. Velocities of less

than 16mm/year are not always without the need for remedial actions eg.: pipelines can be damaged, stabilising
measures can be over-loaded and instrumentation may reach its limit of movement.

TABLE 1 : TERMS DESCRIBING THE VELOCITY OF A LANDSLIDE


(CRUDEN AND VARNES, 1996)
Velocity class
7
6
5
4
3
2
1

Description
Extremely Rapid
-----------------------Very Rapid
-----------------------Rapid
-----------------------Moderate
-----------------------Slow
-----------------------Very Slow
-----------------------Extremely Slow

Velocity (mm/sec)

Typical velocity

5x103

5 m/sec

5x101

3 m/min

5x10-1

1.8 m/hr

5x10-3

13 m/month

5x10-5

1.6 m/year

5x10-7

16 mm/year

Human response
Nil
Nil
Evacuation
Evacuation
Maintenance
Maintenance
Nil

1.2.3
Landslide Classification
Names of the basic movement mechanisms have been defined by Cruden and Varnes (1996):
Fall is the independent movement of rock or soil fragments through free fall, bouncing or rolling. Each fragment
interacts with the slope in periodic collisions, but there is no significant interaction between fragments (cf. fragmental
fall, Evans and Hungr, 1993).
Topple is a rotational movement of steeply dipping columns or slabs of rock or coherent soil separated by
discontinuities.
Slide is a rotational or translational movement of rock or soil masses, sliding on one or several more or less discrete
displacement discontinuities (surfaces of rupture).
Spread is a movement of a mass of rock or soil dominated by internal distortion. The distortion causes thinning and
lengthening of the landslide mass, thus propagating the leading edge forward.
Flow is a spatially continuous movement with a velocity distribution resembling the flow of a fluid.
While these are simple mechanisms, their identification in the field is often quite difficult. For example, moderately
large rock slides may disintegrate, after moving a certain distance, into a swarm of fragments rolling and bouncing
independently. Spreading is also an ambiguous factor, since some greater or smaller degree of internal distortion and
spreading accompanies nearly all sliding motion. Again, some landslides which produce flow-like deposits (earth
flows) in fact move predominantly by sliding on discrete shear surfaces (Hutchinson, 1988). Some flows involve
rheological change in the flowing mass, relative to a rigid base or channel. Others involve continuous distortion of a
homogeneous rock or soil mass (creep movements e.g. Varnes, 1978). Still others may involve flow-like distributed
deformation of a thin fluid zone at the base of a rigid plug. In some cases it is very difficult to determine the precise
movement mechanism, without detailed subsurface measurements. Thus, terms based strictly on movement mechanism
frequently suffer from ambiguities.
Cruden and Varnes (1996) suggested a landslide classification based on a taxonomic order. Landslide names are to
be built from terms describing the following attributes, in sequential order:
- state (Table 3a)
- distribution (Table 3a)
- style (Table 3a)
- rate of movement (Tables 1 and 3(b), (c))
- water content (Tables 3(b), (c))
- type of material (Tables 3(b), (c))
- type of movement (Table 2)
Because both type of movement and rate can change during motion, the same description should be applied
repeatedly to several phases of motion, if necessary. This process is rational and well-organized, but results in rather
long names; for example complex, extremely rapid, dry rock fall debris flow. Further, it is hindered by the
ambiguities of assigning accurate descriptions of movement type, material and water content, some of which have been
discussed above. For some situations therefore it may be sufficient to use the abbreviated classification shown in Table
2 or the simplified version of the Varnes classification shown in Table 4. In other situations it may be necessary to
include a more complete description of the soil material from the mechanical point of view, as done by Morgenstern
(1992) and Leroueil et al (1996).

TABLE 2 : ABBREVIATED CLASSIFICATION OF SLOPE MOVEMENTS


Type of
Movement

Bedrock

Fall
Topple
Slide
Spread
Flow

Rock fall
Rock topple
Rock slide
Rock spread
Rock flow

TYPE OF MATERIAL
Engineering Soils
Predominantly Coarse
Predominantly Fine
Earth fall
Debris fall
Earth topple
Debris topple
Earth slide
Debris slide
Earth spread
Debris spread
Earth flow
Debris flow

TABLE 3 : GLOSSARY FOR FORMING NAMES OF LANDSLIDES


(a)

Activity
State
Active
Reactivated
Suspended
Inactive
Dormant
Abandoned
Stabilised
Relict

Distribution
Advancing
Retrogressive
Widening
Enlarging
Confined
Diminishing
Moving

Style
Complex
Composite
Multiple
Successive
Single

(b) Description on first movement


Rate
Water Content
Material
Type
Fall
Rock
Dry
Extremely rapid
Topple
Soil
Moist
Very rapid
Slide
Earth
Wet
Rapid
Spread
Debris
Very wet
Moderate
Flow
Slow
Very slow
Extremely slow
Note: Subsequent movements may be described by repeating the above descriptors as many times as
necessary

TABLE 4 : A SIMPLIFIED VERSION OF THE VARNES LANDSLIDE CLASSIFICATION. THE


ITALICIZED NAMES REPRESENT LANDSLIDE TYPES LIKELY TO EXHIBIT EXTREMELY RAPID
VELOCITIES (MORE THAN 5 M/SEC). BASED ON FIG. 3.1 IN VARNES (1978).
BEDROCK
FALLS
TOPPLES
SLIDES
SPREADS
FLOWS

COMPLEX

ROCK FALL
BLOCK TOPPLE
FLEXURAL TOPPLE
ROCK SLUMP
ROCK SLIDE
ROCK SPREAD
ROCK CREEP
SLOPE SAGGING

DEBRIS
(<80% sand and finer)
DEBRIS FALL
-

EARTH
(>80% sand and finer)
EARTH FALL
BLOCK TOPPLE

DEBRIS SLIDE

EARTH SLUMP
EARTH SLIDE
EARTH SPREAD
WET SAND AND SILT
FLOW
RAPID EARTH FLOW
LOESS FLOW
DRY SAND FLOW
EARTH FLOW

DEBRIS FLOW
DEBRIS AVALANCHE
SOIL CREEP
SOLIFLUCTION

ROCK AVALANCHE
EARTH SLUMP-EARTHFLOW

Geotechnical engineers working with soils have been more restrained in interpreting movement mechanisms. Until
fairly recent time, geotechnical analysis could be applied only to rigid sliding motion (limit equilibrium methods) or
small-strain deformation analysis of continua. Therefore, sliding was often assumed to be the dominant failure
mechanism. The existence of flow was recognized in the term flow slide, which was defined by Hutchinson (1992)
as a landslide involving the generation of high excess pore-pressures through undrained structural collapse. It is
important to distinguish flow slides, in which full or partial liquefaction occurs on the rupture surface at the point of
failure, from other flows, which fluidize or liquefy in course of motion along the path.

2.

THE MECHANICS OF THE SLIDING PROCESS


As depicted in Figure 5, four different stages of slope movements can be considered:

Post-failure

Pre-failure

Displacement rate

First
failure

Occasional
reactivation
Active
landslide

Time

Figure 5 : Different Stages of Slope Movements (Leroueil et al 1996)


-

The pre failure stage, when the slope is strained throughout, but is essentially intact.
The onset of failure characterized by the formation of a continuous surface of rupture (eg. a shear band)
through the slope.
- The post failure stage, which includes movement of the material in the landslide from just after failure until it
essentially stops.
- The reactivation stage, when the slope slides along one or several pre existing shear surfaces: this
reactivation can be occasional or continuous with seasonal (or longer period) variations in the rate of
movement.
It is recognised that the definition of failure used above is a particular case, and that a slope may have ceased to be
acceptable in the pre failure situation (eg. large deformations may have damaged structures) or in a post failure situation
eg. where sustained creep is acceptable.
2.1
Pre Failure and at Failure Mechanics of Soil at the Micro scale
2.1.1
Some General Concepts
(a) Drained conditions
Figure 6 shows diagrammatic stress displacement curves under drained loading conditions and constant normal
stress, for saturated soils with a low and high clay fraction (finer than 0.002mm) (Skempton 1985).
The weakening from peak strength to critical state strength is due to dilation and softening, and then particle reorientation (alignment of the clay particles), leads to a further reduction until a minimum strength, the residual strength,
is achieved at large displacements. Skempton (1985) adopts the use of the term field residual strength ('RF) as the
strength of fully developed shear or slide surfaces in nature. This value may differ from the laboratory residual strength
depending on the testing method. Skempton (1985), Lupini et al (1981), Chandler (1984), Mesri and Cepeda-Diaz
(1986) and Hawkins and Privett (1985) describe these effects.
Lupini, et al (1981) carried out a series of ring shear tests on sand-bentonite mixtures, and suggested that there were
different mechanisms of residual plane development depending on the clay fraction percentage present. They called the
mechanisms turbulent and rolling shear where the presence of rounded silt sized particles prevents alignment of clay
particles on the shear surface; sliding shear, when the effects of the silt is over-ridden by the predominance of clay
particles; and a transition between these two conditions. These results were summarized by Skempton (1985) and are
reproduced in Figure 7.
These figures show that sliding shear with complete reorientation of the clay particles is likely to occur only
where the clay fraction (finer than .0002mm) exceeds 50% of the total soil, and that for less than about 25% clay
fraction, turbulent or rolling shear occurs without the influence of clay particle alignment.

Figure 6 : Diagrammatic stress displacement curves at constant normal stress 'n (Skempton, 1985)

Figure 7 : Ring shear tests on sand-bentonite mixtures


(from Skempton 1985) after Lupini, Skinner and Vaughan, 1981)
Skempton (1985) presented results of field residual and ring shear tests on a range of soils (Figure 8).

Figure 8 : Field residual and ring shear tests on soils, kaolin and bentonite (Skempton 1985)

Figures 9 and 10 present some other useful information relating 'R and (' 'R ) to the liquid limit.

Figure 9 : Residual friction angle liquid limit relationship (Mesri and Cepeda Diaz 1986)

Figure 10 : Difference between secant effective fully softened


and residual friction angles (Stark and Eid, 1997)
Stark and Eid (1997) attribute the large difference between fully softened and residual strengths for clays of liquid
limits 80-160 (ie. very plastic soils) to the fact that a large displacement is required to realign the particles and not
much realignment has occurred at the fully softened state). For low plasticity soils, the particles are mostly rounded so
there is little difference between the fully softened and residual strength (as shown in Figure 7).
Figures 8, 9 and 10 are useful but should be used with caution, accounting for pre-existing shear surfaces, fissuring,
the clay mineralogy, degree of over-consolidation and any laboratory shear strength data.
Apart from the clay fraction, the mineralogy of the clay sized fraction (% passing 0.002mm) also has an effect on
residual strength. This is particularly so when the clay fraction is large. This reflects the fact that the different clay
minerals have different particle shape and different interparticle bonding. Most clay minerals, eg. kaolinite, illite,
chlorite and montmorillonite are platy structures, and are therefore subject to alignment when sheared. Montmorillonite
has a particularly thin plate structure, and weak inter-plate bonds and leads to the lowest residual strength as low as
'R = 5 o . Kaolinite has a 'R = 15 o and illite approximately 10o.
Some clay minerals do not have a plate structure eg. halloysite has a tubular structure, attapulgite a needlelike
structure, and some minerals with clay sized particles such as gibbsite, haematite and bauxite have essentially granular
structures. This leads to much higher residual friction angles, commonly greater than 25o (Skempton 1985).
Reorientation necessitates particles that are platy in shape, which is mostly associated to the mineralogy of soil
particles, and it is facilitated as the percentage of platy particles increases. This was clearly explained by Lupini et al.
(1981) and justifies relationships between 'R and the plasticity index that reflects to some extent these two factors (e.g.
Voight, 1973). It also appears that for many clayey soils, the relation between residual strength and normal effective
stress is not linear with f/'n ratios decreasing as the applied normal stress, 'n increases (Skempton, 1985; Picarelli,
1991; Stark & Eid, 1997). Two other factors influence the residual shear strength: a) The rate of displacement that

typically increases 'R by 1 to 3 % when it increases by one order of magnitude. This has, however, no significant
influence on the evaluation of the strength to consider for stability analysis (Skempton, 1985; Leroueil, 2000). b) a
factor which can be much more important and is most often neglected is the pore water chemistry. Di Maio (1996a and
b) showed on several clays from southern Italy and a commercial bentonite a change in 'R from about 14 when
prepared with saturated NaCl solution to about 6 when prepared with distilled water. Figure 11 shows this change for
the bentonite. From a practical point of view, this means that the residual strength parameters should be evaluated with
pore water with a chemistry close to the one existing in situ.

Figure 11 : Residual friction angle of Ponza bentonite under normal stresses of 250-400 kPa
for various concentrations of NaCl solution. (Leroueil 2000 after Di Maio 1996(b)
(b) Undrained loading
Figure 12 shows a comparison of effective stress and undrained strength analysis for evaluating the stability of an
embankment. The soil element represented by point X has a mobilised shear stress m. For a contractive soil, the
strength is represented by the undrained strength analysis (USA) with an available strength . ff = Cu (at point Y).

Figure 12 : Comparison of effective stress and undrained strength analysis (adapted from Ladd, 1991)
Point Z represents the strength which is assumed in a conventional effective stress analysis (the effective normal
stress at failure = existing effective normal stress). Pore pressures generated during shearing are assumed to be
dissipated. Point W would apply to an USA for a soil which was dilative.
It is important to note that the strength available in the case of undrained shear of the contractive soil (Cu) is less
than that assumed for conventional effective stress analysis, because of the positive pore pressures generated during
shearing. As discussed below, this can happen in coarse grained as well as fine grained soils.

Figure 13 shows stress strain and stress paths of saturated loose cohesionless sand loaded in triaxial compression
under undrained conditions at a constant rate of strain, undrained creep, and an initial shear stress and cycle loading. In
all cases the large strain or ultimate strength, also known as the steady state strength, or residual undrained strength is
reached at large strains. In an e - log p' diagram (Figure 14(a)), where e is the void ratio and p' is the mean effective
stress, such a test would have its initial point at I1, and the ultimate critical or steady state at point u1 on the steady state
line. For initial condition at point I2, below the steady state line, the undrained behavior would be dilatant with a steady
state at point u2. The position of the initial conditions compared to the steady state line can be characterized by the
state parameter , the difference in void ratio between the initial void ratio and the void ratio on the steady state line
at the same p value (Been and Jeffries 1985).

Figure 13 : Typical undrained response of loose sands in monotonic,


creep and cycle loading
On the steady state line the ultimate shear strength Su is related to the mean effective stress p by the critical state
friction angle. Su can thus be directly related to the void ratio as shown on Figure 14(b). However it is not simple in
practice because the soil specimens should be undisturbed, with the in-situ structure preserved (Poulos et al 1985) and
should be tested in the range of stresses existing in situ. Moreover a small change in void ratio has a significant effect
on the inferred strength, and the steady state line can be influenced by the type of test and stress path used to make the
measurements (Vaid et al 1990).

Figure 14 : Steady state or critical state concept


As discussed in Leroueil (2000), many slopes experience a decrease in effective mean stress with an essentially
constant shear stress, as pore pressures in the slope rise (Stress path I-Y on figure 15(a)). Tests carried out by Anderson
and Riemer (1995), Anderson and Sitar (1995), and Santos et al (1996), behaved in a manner different to that described
above, with some clayey soils consolidated to above the steady state line, showing dilatant behaviour; the same soil
tested in anisotropically consolidated undrained compression tests displayed contractant behaviour. This behaviour can
be explained as follows: (See Figure 15(b)). If initial stress conditions are at point I, on Figs. 15a and b, and if the mean
effective stress is decreased, there is first some swelling of the soil, and yielding or failure is reached at a point such as
Y. Note that the effect of an increase in pore pressure can be different in collapsible unsaturated soils. This latter point
can be on the steady state or critical state line (CSL), as shown on Fig. 15a, or above in structured soils, or below, on the
collapse surface, as observed in loose sands (Sladen et al., 1985; Sasitharan et al., 1993; Konrad, 1993; Lade, 1993). If
initial conditions are at point I1, well above the CSL, the soil at yielding, Y1, is above the CSL (Fig. 15b); for undrained
post-failure conditions, the ultimate strength, in C1, is smaller than the initially applied shear stress, and failure is
followed by an undrained collapse and flow of the soil mass. If initial conditions are at point I3, below the CSL,
yielding would appear at Y3 (Fig. 15b). the ultimate state in undrained conditions, C3, then corresponds to a strength

larger than the shear stress at failure and is thus stabilizing. For the intermediate case with initial stress conditions, I2,
above the CSL and yielding, Y2, below (Fig. 15b), the soil would be dilatant at failure with a tendency to swell to C4.
This seems to correspond to observations made by Anderson & Riemer (1995) on colluvial soil and by Santos et al
(1996) on residual soil. Finally, we can imagine initial stress conditions such as I4 leading to yielding at point Y4,
identical to C4, on the CSL; in this case, the soil would not have tendency to change its void ratio after reaching failure.

Figure 15 : State paths followed in consolidated constant shear drained tests (schematic) Leroueil 2000)
Where the in-situ shear stresses exceed the ultimate strength, flow sliding is likely to occur. Where the shear
stresses greatly exceed the ultimate strength, large, rapid slides, may result. The degree of strain weakening has been
characterized in terms of the brittleness Index IB which has been defined by Bishop (1967):
IB =
Where

p r
p

,%

(1)

p and r are the peak and residual strengths defined under the same effective normal stress.

However, as indicated by Vaughan & Hamza (1977) and by Chandler (1984) the brittleness index alone is not
sufficient to characterise the susceptibility of a soil to progressive failure and the rate at which the strength decreases
from peak strength to ultimate strength is also important. DElia et al. (1998) propose a generalised brittleness index,
IBG, defined as follows:
IBG =
in which

p mob
,%
P

(2)

mob is the mobilised shear stress at the considered strain or displacement. IBG thus varies with strain or

displacement from 0 at the peak to a value equal to IB at large displacements (Fig. 16). Also, in the context of slopes,
IBG, must be associated to stress paths that are representative of those followed in situ, and must thus not be seen as a
fundamental characteristic of a soil. With this extended definition, not only overconsolidated clays, clay shales,
sensitive clays, residual soils and loess, but also cohesionless soils such as loose sands may behave in a brittle manner
in undrained conditions (as suggested by Sladen et al., 1985).
In term of generalised brittleness index, IBG, (undrained behaviour following failure at point Y could be dilatant,
contractant or perfectly ductile, depending on the state path that can be Y3C3, Y1C1 or Y4C4 respectively (Fig. 15b).

Shear stress

IGB =

p - ( or )
p

IGB = IB =

p - r
p

1
Strain or displacement

Figure 16 : Definition of Brittleness Index (DElia et al 1998).


In some contractive soils, eg loose sands, loaded in saturated, undrained conditions, the triaxial compression stress
path is more complex, as shown in Figure 17. The soil reaches a peak undrained strength at A, on an envelope lower
than the strength envelope of the soil.

q1 =

1
1-

1
3

ME
LU PE
VO ELO
T
V
N EN
E
TA
FA C
NS TH
UR
CO ENG
ES
S
R
P
ST
LLA
CO

Su

S ucs

O
B

+
p1 = 1 3
2

Figure 17 : Stress path for saturated loose sand which displays collapse behaviour
The line through these points is termed the collapse surface (Sladen et al 1985, Dawson et al 1998), or critical stress
ratio (Vaid and Eliadorani 1998) or the instability line (Lade 1992). At these stress conditions the soil wishes to
contract, but being constrained by the no volume change of saturated undrained conditions, develops positive pore
pressures resulting in weakening to the ultimate strength Su at B.
The important point here is that the collapse surface strength may be lower than the peak, or critical state (constant
volume) strength, which is commonly used in stability analysis, so failure may occur at stresses lower than expected
(eg: Dawson et al 1998). As pointed out by Sladen et al (1985) for very loose sands whose state lies on the collapse
surface, only minimal excess pore pressure is necessary to trigger collapse. Loading may be essentially drained up to
the point of reaching the collapse surface.
2.1.2
Some situations where contractant and dilatant; drained and undrained conditions occur
Table 5 lists some situations where contractant and dilatant conditions occur, and some references giving examples
Hutchinson (1995) gives a number of examples of contractant, or potential flow conditions and some details
Whether soils behave in a drained or undrained manner depends on its coefficient of consolidation and its
permeability, and the rate of application of the load. What is apparent is that there are many situations where undrained
conditions may occur even in granular materials. These include under earthquake loading, where the cyclic nature of
the applied shear stress generates pore pressures in the few seconds of the earthquake, and in situations where a collapse
mechanism may occur. In these cases, the initial movements may be slow, and in drained conditions, but when the
collapse condition is reached, the rate of shearing is rapid and occurs in seconds, (eg. Dawson et al 1998 description for
mine spoil piles) so even reasonably coarse sands and sandy gravels may behave undrained.
In dams, it is apparent that the loading application of the reservoir may lead to undrained loading conditions. This
has long been recognized for drawdown conditions eg. Duncan (1996), but can occur in some conditions for the
downstream slope (Fell, 2000). It has been well recognized for some time that pore pressures in cut slopes, and under

embankments on soft clays may take many years to reach equilibrium (Leroueil 2000, Skempton 1977, Chandler
1984(a)), and that the critical condition for embankments on soft clays is during construction because the pore pressures
induced by construction dissipate with time, leading to an increase in strength with time (Bishop and Bjerrum 1960,
Ladd 1991).

TABLE 5 : SOME SITUATIONS WHERE CONTRACTANT


AND DILATANT SOIL CONDITIONS OCCUR
Soil Condition
Contractant
(a) Cohesionless

Situation

Submarine and subaqueous slope


Mine tailings
Dredged fills, including hydraulic fill
dams

(b) Cohesive

References for Examples

Silvis and de Groot (1995)


Blight (1997), Dobry and Alvarez
(1967), Troncoso et al (1993)
Sladen et al (1985), Sladen and
Hewitt, (1989), Lee et al (1999)

End dumped mine waste

Dawson et al (1998), Bishop (1973),


Eckersley (1985), Hutchinson (1992).

Uncompacted or poorly compacted road


and rail fills

Davies
and
Christie
Morgenstern (1978)

Loose colluvium on hillslopes

Lee et al (1988), Leroueil (2000),


Fleming et al (1989), Anderson and
Sitar (1995)

Loose dumped dirty rockfill in dams


Normally consolidated, and lightly
overconsolidated clays eg. quick clays of
Canada and Scandinavia

No examples known
Ladd (1991), Trak and Lacasse
(1996), Mitchell and Markell (1977)

Poorly compacted clay fills, puddle core


dams, compacted clays at high confining
stresses

Cooper et al (1997),
Fell (2000)

(1996),

Dilatant
(a) Cohesionless

(b) Cohesive

Dense sandy and gravelly soils eg.


compacted fills compacted rockfill eg. in
dams
Heavily overconsolidated natural clays.
Compacted clay soils at low and
moderate confining stresses

Skempton (1985), Lupini et al (1981)

2.1.3
The behavior of contractant clays
As described in Section 2.1.1, contractant clays may develop positive pore pressures on shearing under many
situations, and the normal design procedure is to use undrained shear strengths, obtained by in-situ (vane shear, cone
penetration test) and laboratory testing by the recompression technique (Bjerrum 1973, Jamiolkowski et al 1985) or
SHANSEP method (Ladd and Foott 1974, Ladd et al 1977, Jamiolkowski et al 1985). The latter method is
cumbersome, requiring the development of an over-consolidation ratio (OCR) versus undrained strength relationship for
each soil in the foundation, and using the OCR of the soil, as measured by laboratory tests, to determine the strength
from this relationship. This last stage of the process is very dependent on good sampling so the pre-consolidation
pressure of the soil can be accurately determined in the laboratory. Several limitations to the SHANSEP technique have
been pointed out, in particular by Leroueil and Jamiolkowski (1991). One aspect is that it does take into account of the
structure of the intact soil. For this reason Ladd (1991) specified that the SHANSEP technique is strictly applicable
only to mechanically over-consolidated and truly normally consolidated soils exhibiting normalised behaviour.
Effective stress analysis methods may be used, but then the pore pressure response to shearing, taking account of the
stress path, must be modelled. As has been recognized for a long time, Jamiolkowski et al (1985), Ladd (1991) and
Kulhawy (1992), the undrained shear strength used in the analysis should take account of the failure mechanism. As

shown in Figure 18, the loading conditions under an embankment are best simulated by triaxial compression, while
beyond the toe, they are best simulated by triaxial extension and direct simple shear.

Figure 18 : Relevance of Laboratory Shear Tests to Shear Strength in the Field (Kulhawy 1992)
Figure 19 shows the mean normalized undrained strength ratios for the major laboratory tests.

Figure 19 : Mean Normalised Undrained Strength Ratio for Major Laboratory Tests (Kulhawy 1992)
For this figure the reference strength is given by the modified Cam Clay model, which Kulhawy (1992) indicates is
reasonable for relatively unstructured soils. For sensitive, cemented and other structured fine grained soils, the
reference strength is a lower bound. It can be seen that to use Ko triaxial compression tests is unconservative for those
parts of the failure surface best modelled by triaxial extension and direct simple shear tests.
For much routine practice, the strength is estimated using vane shear or cone penetration tests, modified by
empirical correction factors eg. Bjerrum (1973), Lunne et al (1976), Azzouz et al (1983), Aas et al (1986) and Hanzawa
(1989). These methods have the advantage of providing a continuous or near continuous strength profile. The
correction factors account for anisotropy and rate effects (and Azzouz et al (1983) for 3 dimensional effects). They are
developed by back analysis of failures and are essentially empirical, and de-facto allow for the stress anisotropy
discussed above. Ferkh and Fell (1994) re-analyzed these case studies making consistent assumptions in regard to the
other major variables which come into analysis of the stability of embankments on soft clay, ie the allowance for
fissuring and cracking in the over-consolidated crust of the soil, cracking (or the lack of it) in the embankment, the
presence of shells and roots affecting in-situ tests, and the variability of the strengths measured by the in-situ and
laboratory tests.
It became apparent that these factors had a major effect on the back-analyzed factors of safety (up to 0.3 or 0.5). If
fissuring is present, vane shear and CPT or CPTU tests over-estimate the strength. The reported presence of fissuring at
depth seemed to be related to soils with a liquidity index less than 1.2, and over-consolidation ratio of 1.8 or more.
However the best method to check for fissures is to take undisturbed tube samples, air dry and break them for
inspection. Corrections have been proposed by a number of investigators. For example, Tavenas and Leroueil (1980)
and Lefebvre et al. (1987) suggested that the operational strength in the crust is given as shown in figure 20. Ferkh and
Fell (1994) give an overall design method. Lefebvre et al (1987) describe methods to estimate the strength of the overconsolidated crust using plate load tests.

Figure 20 Suggested methods for estimating the strength of the over-consolidated crust
For embankments constructed in several stages, experience in eastern Canada shows that the measured vane shear
strength does not need to be corrected (Tavenas et al., 1978; Law, 1985). One consequence is that the strength increase
under an embankment cannot be directly evaluated from the measured vane strength values. On the other hand, Demers
(2000) shows, that the cone penetration test can be used for the evaluation of the strength increase under stageconstructed embankments, estimated directly from the increase in net tip resistance (qT - VO).
2.1.4
The behaviour of contractant granular soils
The susceptibility of granular soils to collapse, or static liquefaction, is dependent on the particle size distribution,
void ratio (or density index), the initial stress conditions, and the stress path of loading eg. triaxial compression or
extension. This has been an active research topic in recent years, and the following summarizes some of the recent
findings.
Lade and Yamamuro (1997), and Yamamuro and Lade (1997), show that there are four different types of undrained
stress paths after the instability line (collapse surface). These are shown in Figure 21:

Static liquefaction occurs at low stresses and is characteristed by large pore pressure development resulting in
zero effective stresses at low axial strain levels (stress path AO). Increasing confining pressures result in
increasing effective stress friction angles in this stress region.

Temporary liquefaction occurs at higher stresses than the static liquefaction region, and is characterized by
stress path BCDE. Hence the behavior is contractant from C to D, then dilatant from D to E which occurs at
large axial strains. The tendency to dilate (and less likely to contract) increases as the initial confining stresses
increase.

Temporary instability (stress path FGHI) is similar to temporary liquefaction except that the amount the stress
difference increases beyond the initial peak (ie I vs G), is not as large as that exhibition by temporary
liquefaction.

H
E

C
D
A
O

Figure 21 : Undrained effective stress of loose silty sands (Lade and Yamamuro (1997))
The mechanism of volumetric contraction for static liquefaction and temporary liquefaction are related to particle
realignment, particle crushing and subsequent rearrangement of voids for temporary instability and instability. The
mechanism of volumetric contraction can also be due to redistribution of voids, and hence void ratio in the specimen.
Yamamuro and Lade (1997) argue that the reported cases of static liquefaction occur at low confining stresses (most

slopes under 20m high) consistent with the static liquefaction behavior. Others eg. Stoutjesdejk et al (1998) and
Dawson et al (1998) indicate that flow slides can result in the temporary liquefaction region. This would seem to relate
to disagregation of the soil in the displacements occurring in temporary liquefaction. Lade and Yamamuro (1997)
tested sands with varying silty fines content and showed that at low confining stress (25kPa) the potential for static
liquefaction was increased with increasing fines content. Clean sand exhibiting static liquefaction at a relative density
of 20%, while soil at 50% fines content exhibited static liquefaction with relative densities up to 60%. They explained
this in terms of the distribution of fine and coarse grains. They acknowledge the results are contrary to earthquake
liquefaction observations, but point out it may be applicable to hydraulic fills, or very young natural soil deposits.
Vaid, and his co-workers (Uthayakumar and Vaid (1998), Sivathayalan and Vaid (1998), Vaid and Eliadorani
(1998)) have shown that: The undrained response of loose sand is highly dependent on the load direction ie whether
triaxial compression ( = 0o and b = 0) and triaxial extension ( = 90o, b = 1) where = major principal stress
direction with respect to vertical, and b = intermediate stress parameters (2 - 3)/( 1 - 3).
Figure 22 shows the dependence of brittleness index IB on , friction angle mobilized at peak deviator stress (ie. on
the collapse surface) on , and the dependence of peak shear strength on for two sands. In these diagrams 'mc is
the effective mean normal stress at the end of consolidation.

(a) Dependence of brittleness index IB on .

(c) Dependence of friction angle mobilized


at peak deviator stress

Figure 22 : Effect of loading direction on undrained


behavior of two sands (a) Brittleness Index IB versus
(b) friction angle mobilized at peak deviator stress versus
(c) peak shear strength versus (Uthayakmar and
Vaid 1998)

(c) Dependence of peak shear strength on


It can be seen that samples tested in triaxial extension are more brittle and mobilize a lower peak strength (ie the
collapse surface is lower and the peak undrained strength is lower) than those tested on triaxial compression. The
ultimate (or in some cases quasi-ultimate strength (ie at point B in Figure 21) is also lower in triaxial extension Figure
23.
These results indicate that the use of triaxial compression tests is unconservative. Soils which are dilative under
triaxial compression may become contractive under triaxial extension.
The ultimate strength is not uniquely related to void ratio alone, but on the direction of the principal stresses, the
effective mean normal stress prior to shearing, and the magnitude of the intermediate principal stress.

(a) Dependence of quasi steady state or steady state strength


on for Fraser River sand

(b) Dependence of quasi steady state or steady state


strength on for Syncrude sand

Figure 23 : Effect of loading direction on ultimate or quasi ultimate strength for two sands
(a) Fraser river sand (b) Syncrude sand (Uthayakumar and Vaid 1998)
An extreme example of the effect of anisotropy on the stability of slopes is provided by dredging works associated
to the construction of the Bangabandhu Bridge in Bangladesh (Hight et al., 1998). Trenches excavated below water
appeared unstable, with generation of flow slides, at slopes of 1 to 5 (slope angle of 11 o). The material in which the
slopes were dredged were micaceous sands, with a mica content lying between 5 and 10%. The relative density was
generally in excess of 50%, a value which, according to Hight et al. (1998), would not normally be associated with a
potential of flow. Hight et al. (1998) attributed this behaviour to the collapse potential and the extreme anisotropy
induced by the presence of a small amount of mica. This can be illustrated by the following selected results: while the
clean sand in constant volume simple shear tests shows a highly dilatant behaviour, the same sand with only 1% mica
becomes brittle and show a potential to collapse; triaxial undrained compression and extension tests performed on the
natural material, prepared at relative densities of 58% and 55% relatively gave undrained strengths in excess of 180 kPa
in compression and of only 10 kPa in extension.
Vaid and Eliadorani (1998) showed that contrary to normal understanding, partially drained states could be more
prone to instability than totally undrained. Sivathayalan and Vaid (1998) showed that membrane penetration in testing
can affect drainage state and potential for liquefaction.
de Groot and his co-workers (Silvis and de Groot 1995, de Groot and Stoutjesdisk 1997, Stoutjesdijk et al 1998)
describe methods for predicting flow slides in underwater slopes. Their work seems to indicate that for the apparently
clean sands tested, collapse behavior occurred only for a density index less than 30% ie very loose.
Dawson, et al. (1998) describe liquefaction flow slides in coal mine waste dumps. These occur in dumps higher
than 70-100m, loose dumped at the angle of repose, onto hillsides with slopes less than 18o. Most failed during
operation but failures occurred up to 13 months after placement ceased. Failures usually occurred in times of rain or
snowmelt.
The grain size distribution of the dump materials are shown in Figure 24(a) and in 24(b) along with other flow
liquefaction cases (including Aberfan). Laboratory testing showed a tendency for collapse behavior at void ratios
greater than 0.3 with a peak strength (collapse surface) of 28o, compared to the angle of repose of 38o. These strengths
are consistent with the factor of safety of 1.0 at the initiation of failure. Dawson et al. found that crushing during
consolidation appears to reduce the brittleness at void ratios less than 0.3 by breaking down the collapsible structure.
The brittleness index varied from 0 at a void ratio of 0.3 to 15-45% at a void ratio of 0.4. The ratio of the ultimate
strength (steady state undrained strength) to the vertical stress (for Ko consolidation) was around 0.15-2, and gave a
factor of safety 0.7. Hence in these cases, even a relatively modest amount of strength loss, produced rapid debris
flows.
Law et al (1998) describe some laboratory tests on loose remoulded granitic soil from Hong Kong. These soils
which had fines content of 40% including 12% - 20% clay, exhibited collapse tendencies when tested undrained at
void ratios from 0.64 to 0.78 (corresponding to a dry density ratio range of about 75% to 85% of standard maximum dry
density). The slope of the collapse surface appeared to be dependent on the void ratio. In related work, HKIE (1998)
found that the collapse surface could be modelled as c = 3kPa, ' = 20o for fill slopes with a dry density ratio greater
than 75%, compared to a steady state friction angle of 36o to 40o. Pre-1977 fill slopes, which were generally uncompacted had void ratios generally in the range 0.5 to 1.5; dry density ratio usually 70% to 85%. Failures of fills in

Hong Kong have exhibited flow, including those at Sau Mau Ping in 1972 (Government of Hong Kong, 1972). Yim
and Siu (1998) report that studies by Binnie and Partners (1982) demonstrated that fills would be liquefaction prone if
the dry density ratio was less than about 88% for decomposed granite, and 93% to 97% for decomposed volcanic (silty
clay).

(a) The Canadian Mine Waste

(b) Sandy gravels associated with liquefaction behaviour

Figure 24 : Grain size distribution for mine waste which have experienced flow failures (Dawson et al 1998)
Overall this work tends to raise more questions than it answers, and it is difficult to give clear advice on what is
reasonable in a practicing as opposed to research environment, but we present the following as a summary.
(a) Susceptible Soils

Sands with a density index of less than about 30% may be susceptible to collapse.

Silty sands may be more susceptible to collapse than sands, and may be susceptible to collapse at higher
density index.

Mine waste materials, which classify as sandy gravels, with as little as 3% silt, may be susceptible to collapse
if the void ratio is more than about 0.3.

Poorly compacted, or loose dumped (void ratio 0.63-0.78) completely decomposed granite fill from Hong
Kong with up to 40% silt and clay fines is also potentially subject to collapse.
These are only examples and should not be relied upon as definitive classes of collapse susceptible soil. Laboratory
tests should be carried out for each case.
(b) Testing methods
Tests to determine the collapse surface, and ultimate strength should be carried out to model the load condition and
stress level. This is likely to include triaxial extension as well as triaxial compression tests, carefully following the
stress paths expected in the field.

(c) It is clear that prevention is the best and most straight forward cure in some situations, eg. it is best to compact
fills so they will be dilative, and even in some situations of existing structures, it will be best to compact loose material
in cuts and fills, or at least the outer part of them (and provide drainage under the compacted fill), rather than attempting
to analyze whether they will or will not be contractive.
2.1.5
Other strain weakening situations high clay content soils
First time slides in high clay content (% passing 0.002mm) soils are strain weakening due to particle re-orientation
on the surface of rupture and to loss of the effects of bonding in the miscrostructure of the soil . If the soils are overconsolidated, there is a further component of weakening due to dilation (Figures 6, 7 and 8).
The degree of strain weakening in drained loading can be judged by calculating the brittleness index. Figure 8, 9
and 10 are useful in assessing this in the absence of shear strength data, and as a check on laboratory or back-calculated
estimates of 'R . However it must be recognised that in undrained loading the brittleness will be affected by the stress
path as discussed above.
2.1.6
Unsaturated soil shear strength
In general, unsaturated soils have a shear strength that is larger than that of the same soil saturated. Fredlund et al.
(1978) proposed that it could be expressed as a linear combination of the net normal stress ( n u a ) and of the matric
suction ( u a u w ) :

f = c' + ( u a u w ) tan b + ( n u a ) tan '

(3)

where:
c'
=
effective cohesion (the same as for the saturated soil)
'
=
effective angle of internal friction (the same as for the saturated soil)
tanb
=
strength parameter associated with matric suction
n
=
total normal stress
uw
=
pore water pressure
ua
=
pore air pressure
(n ua)
=
net normal stress
(ua uw)
=
matric suction
The term (ua uw) tanb appears as a cohesion due to suction. Experimental evidence shows that Equation 3 is valid
for a variety of unsaturated soils, but, most of the time, within a limited ranges of matric suction and net normal stress.
Also, limitations to the use of Equation 3 have been found:
(a) As evidenced by Escario and Saez (1986) and others since, at suctions smaller than the air entry value for the
soil, as long as the soil remains saturated, is equal to . In these conditions, a change in pore pressure is
equivalent to a change in effective stress.
(b) Testing silts, Delage et al. (1987) and Matouk et al (1995) observed strength envelopes obtained at different
suctions which were converging towards a unique point, indicating that there is no effect of suction for net normal stress
in excess of a given value.
(c) It is known that in some soils, cohesionless soils in particular, shear strength first increases and then decreases
as suction is progressively increased. Examples of such behaviour are reported by Escario and Juca (1989) and by
Fredlund et al. (1995b).
b

b
The other limitation of Equation 1 is that is a strongly non-linear function of suction. This limits the predictive

capacity of the approach to situations that the range of suction values used in the laboratory to determine is the same
as that expected in the field (Khalili and Khabbaz, 1998). Also the approach requires extensive and time consuming
laboratory testing in an unsaturated state. The equipment used for this purpose is often beyond that of many
geotechnical engineering laboratories.
A more promising approach appears to be based on the effective stress approach (Bishop, 1959, Kogho et al. 1993,
Modaressi and Abou-Bekr 1994, Fleureau et al 1995, berg and Sllfors 1995, Khalili and Khabbaz 1998, Geiser
2000),
b

= ( u a ) + (u a u w )

(4)

where is the effective stress parameter, which has a value of 1 for saturated soils and 0 for dry soils. The shear
strength f is then evaluated as:

= c + tan

(5)

which is identical to the relationship used for saturated conditions.


In Equation (4), is an empirical parameter representing the proportion of soil suction that contributes to the
effective stress. Several attempts have been made to define this parameter in the past. Many authors have correlated
to the degree of saturation, with no real success. In a more recent approaches Khalili and Khabbaz (1998) considered
as a function of the suction. They proposed a unique relationship between the effective stress parameter and the ratio
of suction over the air entry value, based on the shear strength data of 14 soils reported in the literature. They expressed
as:

(u uw )
= a

(ua uw )e
where

(u a u w )e

0.55

(6)

is the air entry suction.

The advantage of this effective stress approach is that it leads to significant simplifications for practical purposes. It
reduces the model parameters and eliminates the need for shear strength tests in an unsaturated state. In addition, it
enables saturated and unsaturated states to be considered simultaneously, as the transition from saturated to unsaturated
state is treated in a straightforward manner.
2.2

Pre Failure and At failure Mechanisms of Rock


These aspects are covered in other papers in this conference:
J. Read, E. Hoek, Z. Chen and A. Karzulovic (2000), Rock Slopes in Civil and Mining Engineering, and H.
Einstein, G. Mostyn, P. Marinos, O. Stephansson and I. Johnston (2000), Shear Strength of rock Masses and
Discontinuities.
2.3
Pre Failure and Failure Mechanisms in Soil Slopes at the Macro Scale
2.3.1
Creep
Creep in the context of this paper is deformation with time under deviatoric stress. It is a pre-failure phenomenon,
so in slopes, only relates to first time sliding. It is not meant here to describe slow movements of slopes due to shrinkswell and freeze-thaw, or to re-activation of movement of landslides.
There are various ideas on what causes creep:
(a) At the micro scale (soil substance), it has been attributed to the breaking of bonds and the re-orientation of soil
particles (Ter Stepanian 1975); visco-frictional sliding relating to the inter-particle contacts, (Kruhn and Mitchell 1993);
pore pressure increases and weakening of the clay skeleton by a fatigue phenomenon (in undrained creep) in
unstructured clays (Lefebrve and LeBoeuf 1987). There is also evidence that the effects of strain rate and temperature
combine in a unique viscosity law (Boudali et al 1994, see Leroueil and Marques 1996), which could be explained by
the rate process theory related to the activation energy at the level of molecules, atoms or particles. This theory is
synthesised by Mitchell (1993).
(b) At the macro scale, Mitchell (1993) explains creep in over-consolidated clays in undrained loading as shear
deformations cause dilation and development of negative pore pressures, which do not develop uniformly throughout
the sample, but concentrate along planes where the greatest shearing stresses develop. With time during sustained
loading, water migrates into zones of high negative pore pressures leading to softening and strength decrease relative to
the strength in normal undrained strength tests. Figure 25 shows the effect.
As explained by Mitchell, the effective stress path is represented by AB, and AC represents the total stress path in a
conventional CU (consolidated undrained) test. The negative pore water pressure at failure is CB. If a creep stress DE
is applied to the same clay, a negative pore pressure EF is induced. This negative pore pressure dissipates during creep,
and the clay in the shear zone swells. At the end of the creep period, the effective stress will be as represented by point
E. Further shear starting from these conditions leads to strength G, which is less than the original value at B. It is
evident also that if the negative pore water pressure is large enough, and the sustained load is applied long enough, then
point E could reach the failure envelope.
(c) It has been evidenced that the strength envelope of clayey soils in their over-consolidated domain is time or
strain rate dependent (Tavenas et al 1978; Leroueil and Marques, 1996). Figure 26 also provides evidences the fact that
the pre-consolidation pressure is strain rate dependent. Creep failure in this domain thus results from the combined
effect of a lowering of the strength envelope and change in pore pressure. On the other hand, there is no significant
strain rate effect on the normally consolidated friction angle, so that a reduction of undrained shear strength with time or

decreasing strain rate in this domain is associated only to pore pressure increase. These phenomena results in an
undrained shear strength that typically decreases by 10% when strain rate decreases by one order of magnitude (figure
27).
(d) These viscous phenomena are now well recognised for laboratory conditions. However, the links with
practical applications are almost not existent. In particular, the rate of testing is not discussed in these terms and would
certainly be an interesting aspect for research.

Figure 25 : Stress paths for normal undrained shear,


and drained creep of heavily over-consolidated clay (Mitchell 1993).

'3 ) 2 (kPa)

120

Numbers close to the points


140
are strain rates in 10 -8 s -1

140

140
-6

0.2

10

80

23

140

-1
vs (s )

( '1

100

-7

0.6
0.2

10

0.3

-8

10

0.6

-9

10
60
70

3' ) 2 (kPa)

150

( 1'

100

90

110
130
( '1 + '3 ) 2 (kPa)

-6
CAU 1 = 1.4 x 10 s -1
-6 -1
CIU 1 = 1.4 x 10 s
CAD 1 = 2.3 x 10 -7 s -1
CAU 1 = 6.0 x 10 -9 s -1
CA-creep
From one-dimensional
CRS and creep tests
Compression tests at
constant '3 / '1 ratio

150

L
CS

Ko

-6 -1

line

vs = 10 s

vs = 10 -9 s -1
50

50

100

150
200
( '1 + '3 ) 2 (kPa)

250

300

Figure 26: Influence of strain rate on the limit state of Mascouche clay
(Leroueil and Marques, 1996, after Marchand 1982)

1.5

Cu
= 1 %/h
Cu
for
u

Cu
= 1.00 + 0.10 log
C u at 1 %/h

1.0

0.5
-3
10

26 clays

-2

10

-1

10

10
10
Strain rate, (%/h)

10

10

10

Figure 27 : Influence of strain rate on the undrained strength


measured in triaxial compression (Kulhawy and Mayne, 1990)
Eppalock Dam, a central core earth and rockfill dam, with poorly compacted rockfill shoulders, experienced
successively larger vertical and horizontal displacements in three major reservoir drawdowns. Figure 28 shows the
displacements and reservoir levels. In creep terminology, tertiary creep was experienced in each of the drawdowns.
This was investigated and analyzed at some detail (Woodward Clyde (1998), Fell (2000)) and it was shown that the
upper part of upstream slope had developed a rupture surface, and this was due to cracking and softening which had
occurred in the compacted earthfill above the reservoir. When the failure occurred (on the first or subsequent
drawdowns) is not known.

Figure 28 : Eppalock Dam Observed Crest Settlement (Goulburn Murray Water)


Some general characteristics of creep behavior are:
(a) Creep deformation is observed at all levels of deviatoric stress (Singh and Mitchell 1968 (a) and (b), Bishop
1967). Figure 29 shows schematically the stress conditions in a saturated slope (Tavenas and Leroueil, 1981, Leroueil
et al 1996).
In a perfectly stable slope, stress conditions are in the overconsolidated domain, and can be represented by a point
such as D in Figs 29a and 29b. If the slope is subjected to either loading at the crest or erosion at the toe, the stress

conditions move respectively to L or E on Fig. 29a. As the groundwater regime in the slope varies seasonally, the stress
conditions move from Lw corresponding to low water conditions to Hw when high water conditions prevail (Fig.
29(b)). According to the model described by Tavenas et al. (1978), the creep rates also vary with the seasons, being
much higher when water conditions are high. This has been confirmed by field observations.

'1- '3
2

CS

0.1
0.01 E
0.001
0.0001
0.00001

'1- '3
2

CS

L
(Yi

D
O

O: before formation of the slope


'1 +'3
2
(a)

0.1
0.01
Hw
0.001
0.0001
0.00001

(b)

Lw

(Yi
O
'1 +'3
2

(after Tavenas et Leroueil, 1981)

Figure 29 : Schematic creep effects and effective stress conditions in natural slopes
(a) For loading at the crest (DL) and erosion at the toe (DE), (b) for pore pressure fluctuations,
DHW and DLW, (Tavenas and Leroueil, 1981 Leroueil et al 1996,).
Local failure as initiated when the stress state, either E on Fig. 29(a), or Hw on Fig. 29(b), reaches the peak strength
envelope corresponding to the age or the strain rate of the slope. After the peak, stress conditions progressively move
towards the critical state line (CSL), and part of the shear stress is transferred to the neighbouring soil elements which
could then reach local failure. This phenomenon of progressive failure can extend into a continuous shear surface
through the entire soil mass, and then trigger a landslide.
(b) The process of creep has been defined as occurring in 3 stages as shown in Figure 30; primary (at a decreasing
strain rate), secondary (at a constant strain rate), and tertiary (at an accelerating strain rate), generally to creep rupture or
failure. Varnes (1982) and others have noted the short period of secondary creep, and the possibility that it is due to
concurrent processes of primary and tertiary creep.

Figure 30 : Diagrammatic representation of creep for moving slopes


Based on creep tests, Singh and Mitchell (1968(a) and (b)) proposed the relationship below to describe primary
creep for deviatoric stresses 30-90% of the peak strength.


where


A

=
=
=
=

= A e D ( t1 / t ) m

strain rate of creep


parameter which reflects the order of magnitude of the creep rate
slope of the relationship between logarithm of creep rate and stress intensity
stress intensity = D/Dmax

(7)

D
Dmax
t1
t
m

=
=
=
=
=

deviator stress
peak deviator stress
reference time
time
slope of the log strain versus log time plot.

Tavenas et al (1978) confirmed that the equation proposed by Singh and Mitchell (1968(b)) generally describes the
volumetric and shear strain creep behavior of an over-consolidated clay. The power parameter, m, was determined at
various stress levels along a number of stress paths and was generally in the range 0.7 to 0.8. Tian et al (1994) and
Silva & Brandes (1996) verified the equation for drained creep tests on cohesive marine sediments. Tian et al (1994)
commented that the creep parameter, m, is not constant for NC clays, it depends on the applied stress level and
plasticity. Silva & Brandes (1996) commented that the creep parameter, ma, varies with time, has a small dependence
on stress level and there was not much difference between ma for undisturbed and reconstituted samples.
(c) Soils subject to creep will fail at deviatoric stresses less than the peak strength. This has been demonstrated in
laboratory tests by a number of researchers including Lefebvre (1981), Tavenas et al (1978), Murayama and Shibata
(1958). Lefebvre (1981) has shown that the peak strength measured in high quality undisturbed samples of eastern
Canadian clay is much higher than that mobilized in-situ and found that a strength at about 8% strain, well beyond the
peak, was consistent with observed failures.
Leroueil (1998) and Hunter and Khalili (2000), based on the concepts of Lefebvre (1981), and Varnes (1982), and
an assembly of laboratory creep test data, have proposed that:

Creep to failure can occur at less than peak strength with the limiting strength possibly being as low as the
fully softened (or critical state) strength (figure 31(a)).

The level of shear strain at which the onset of failure due to creep occurs is equivalent to the shear strain at
peak stress ( peak ) in the equivalent conventional strength test. The onset of failure is defined as the point of

minimum strain rate, the change from primary to tertiary creep behavior.
The time to onset of failure can be predicted based on the accumulated shear strain (Figure 31(b)).

Figure 31(a) : Idealised creep to failure at


Stresses Less than peak strength (Hunter

Figure 31(b) : Onset of tertiary creep when


accumulated strain peak (Hunter and

and Khalili (2000)

Khalili (2000)

The results of a limited number of sets of drained creep tests on over-consolidated and sensitive structure clays
appear to satisfy the hypothesis. Further testing and assembly of other test data is being undertaken.
Using a simple numerical model based on these concepts, Leroueil (1998, 2000) showed that equation 7 results from
the fact that the soil behaviour is strain rate dependent, and that the stress is applied instantaneously at time zero in
creep tests.
The major gaps in knowledge at this time are the linkage between laboratory and field performance, and the physical
understanding of the creep processes, both at the micro and macro scale. In many situations observed creep behaviour
in the field may be triggered by changes in pore pressure and loading or by softening processes, similar to those
suggested by Mitchell (1993) and observed in Eppalock Dam and it is important to recognise this.
As will be discussed in Section 4.6 , reactivated landslides also exhibit increased rates of deformation as the yield
condition is approached. It may be that these phenoma are not very different to some creep processes.
2.3.2
(a)

Progressive and Retrogressive Failure


Progressive failure

The term progressive failure is used to describe situations where the soil (or rock) is strain weakening, and this
results in areas of high stress in a slope reducing in strength as the soil yields (either in drained or undrained loading)
with the stresses in the slope redistributing to adapt to the changed yield strength. To have progressive failure, it is
necessary to have non-uniformity of shear stresses, and boundary conditions such that strains exceeding failure may
develop. This may progress through to collapse of the slope.
Progressive failure may occur in any strain weakening material such as dense sand, loose sand (in undrained
loading), normally and near normally consolidated clays (in undrained loading) eg. soft marine, and quick clays, heavily
over-consolidated, high clay content clays, and fissured clays, and rock.
Bjerrum (1967) describes a particular situation where progressive failure may occur. This is the case of a cutting or
slope in gently sloping stiff clay or weak, rock with a strain weakening layer as shown in Figure 32. The slope shears
progressively along OP1P2, as the in-situ stresses E are relieved. Translational sliding will occur if the gravity stresses
exceed the residual shear strength.
However it must be recognised that this is not the only situation where progressive failure may occur. Any slope
which has non-uniform stresses (ie virtually all), some of which are at the yield stress, and with strain weakening soil
(or rock) will experience progressive failure.
The process of progressive failure can be described as follows. If shear stresses locally reach the peak shear strength
of the material, there is local failure. If the soil presents some strain-softening behaviour, the failed soil elements will
support a decreasing shear stress as strain increases. The part of the shear stress which is not supported anymore by the
failed elements is then transferred to the neighbouring soil elements that can fail in turn. The process continues until an
equilibrium between shear stresses and strains (or displacements) has been reached. At that time, along a potential
failure surface, part of it can exceed the peak, with possibly some elements at large deformation or residual strength,
whereas another part of the potential surface has not reached the peak. If such equilibrium cannot be obtained, the
process will continue until failure conditions extend along the entire failure surface. This process depends to a large
extent on the brittleness of the soil.

Figure 32 : Progressive failure mechanism in a cutting in strain weakening stiff clay (Bjerrum 1967)

Examples of progressive failure are given in Bjerrum (1967). Well documented case studies include the Saxon Clay
Pit (Burland et al 1977); Edmonton Convention Centre (Morgenstern 1990, Chan and Morgenstern (1987); Carsington
Dam failure (Vaughan 1991, Potts et al 1990); San Luis Dam (Stark and Duncan 1991). The best illustration of
progressive failure in over consolidated clay is provided by the Selborne cutting stability experiment (Cooper, 1996;
Cooper et al, 1998).
Figures 33, 34 and 35 show the geometry of the cut, the location of instruments, inclinometer profiles at different
times and the progressive development of the slide surface.
The first readings were taken shortly after the end of excavation, approximately on day 400. Figure 34 shows
lateral displacement profiles in inclinometers I-04, I-06 and I-08 at selected times. These results as well as observations
reported by Cooper and co-workers give interesting indications on the development of progressive failure:
All the inclinometers show a progressive deformation of the soil mass with occurrence of local shearing at
some time. This is quite clear for inclinometer I.06 which shows continuous deformations on day 88 (reading
C on Figure 34b), initiation of localization of displacement on day 175 (reading D), and clear localization on
day 186 (reading E).
Shortly after excavation, localized shearing appeared at the toe of the cut, whereas the overall factor of safety
was larger than 1.26. As shown in figure 34c, localization was already evident on day 171 in the
inclinometer I.08, at a depth of about 2 m. Figure 35a shows the part of the final slip surface on which failure
was already reached at that time.
At the time of reading C (days 88-96), the profile in I.06 is continuous whereas localizations were observed at
a depth of about 2 m in I.04 and I.08. Figure 35b shows the localizations at the higher and lower parts of the
final slip surface.
At the time of reading D (day 175), localization was discernible in all the inclinometers, except inclinometer
I.05 (Figure 35c). According to Cooper et al. (1998), the global factor of safety was close to 1.04 at that time.
The last reading (reading E on day 186) shows localization along the entire slip surface, and failure occurred 10 days
later (figure 35d). Figure 34a (I.04) also shows that another slip surface is developing, about 4 m below the first one
(Figure 35d).

Figure 33 : Selborne cut, with recharge zone, and inclinometers (Cooper et al 1998)

Figure 34 : Selborne cut inclinometer profiles at different times (Bromhead et al 1998)

Figure 35 Selborne cut Development of the slide surface (Leroueil 2000, after Bromhead et al 1998)
The strain developed in the slope, and along potential failure surfaces is not uniform, so the mobilized shear strength
(or degree of strain weakening between peak and residual) is not constant, and conventional limit equilibrium analysis
using peak strength will over-estimate the factor of safety, and those using residual strengths, will under-estimate the
factor of safety.
A number of authors have developed simplified methods to allow for this in Limit Equilibrium Analysis (LEA).
These are summarized in Mostyn and Small (1987). Of these, the methods by Law and Lumb (1970), Chirapuntu and
Duncan (1977), and Cooper (1988) seem reasonable.
However as pointed out by those authors, the methods are only approximations. To model the slopes correctly,
numerical analysis is necessary. These need to model the strain weakening of the soil, (which is not simple, particularly
if the strain weakening material is not in a confined layer.
However approximate numerical analyses, combined with LEA using the range of possible strengths, can give a
reasonable understanding of the slope behavior.
Numerical analyses of cuts and slopes (see Section 4.4) show that progressive failure may develop not only along a
potential failure surface, but in possibly large volume of the slope, even if the slope remains stable (Kovacevic, 1994;
Leroueil, 2000). According to the latter author, reaching the peak strength means some destructuration of the soil mass
and this can be observed by piezocone profiles carried out below the crest of slopes of precarious stability. Figure 36
presents piezocone profiles performed in a slope of marginal stability. The deposit is constituted of very stiff clay and
silt, but with a liquidity index only slightly smaller than 1.0. It can be seen that the profile carried out 32 m behind the
slope is regular, slightly increasing with depth, while those obtained closer to the slope are very irregular and with
smaller (qT vo) values in their upper part. Such weakened zones were observed in at least 6 different slopes, prior to a
first time failure.
(b) Retrogressive failure
Retrogressive failure refers to the situation where part of a slope fails, and removal of the support from this part of
the slope, results in the failure extending into a region where previously the factor of safety was greater than one.
Figure 37 gives examples

Figure 36 : Piezocone CPTU Profiles at La Baic (Leroueil 2000 after Delisle and Leroueil 2000)

Figure 37 : Example of retrogressive failure


(Leroueil et al 1996, as modified after Poschmann et al 1983)
Tavenas (1984) discusses the situations where retrogressive failure may occur in soft clays. In these situations the
retrogression may be rapid because of the large loss of strength in undrained shear. Based on the work of Tavenas et al
(1983), Mitchell and Markell (1974) and Lebuis et al (1983), Tavenas (1984) and then Leroueil et al (1996) concluded
that the following factors make retrogressive sliding likely:
(i) The ability of the clay to be remoulded assessed by the stability number

Nc = H/Su
Where

Nc

H
Su

=
=
=
=

(8)

stability number
unit weight of the soil in the slope (kN/m3)
Height of the slope (m)
undrained shear strength (kN/m2)

Retrogression may occur for Nc 4 for clays with a plasticity index of 10, and for Nc 7 or 8 with a plasticity
index of 40.
(ii) The ability of the clay to flow away from the toe of the slope when remoulded This will occur when the
remoulded shear strength Sur is less than 1kPa, or if the liquidity index ((water content plastic limit)/plasticity index)
is larger than 1.2.
Trak and Lacasse (1996) discuss their experiences where flow slides in soft clays may occur, they concur with the
above criteria, and give some additional guidance on the likely presence of quick clays.
These include:

Soils of glacial origin, or deposition in a marine environment and post depositional leaching

A preponderance of non clay minerals (quartz and feldspar) in the clay fraction (<0.002mm) or the soil.

Plasticity index less than 20%, high liquidity index (>2).


However caution should be applied in using these criteria in clays other than those for which they were derived ie in
Eastern Canada and Scandinavia.
Retrogressive failure, with flow, is a characteristic of collapse type failures in granular soils. Retrogressive failure is
also a common feature of landslides and failures of cuts and fills, in less brittle soils, even though the failed material
does not flow away from the slope.
2.3.3
The Shear Strength of Fissured Soils
Fissures are discontinuities in soils which have similar implications to the strength and permeability that joints have
in rock masses they reduce the soil mass strength, and (usually) increase the soil mass permeability.
Fissures in soil have a wide range of characteristics relating to the properties of the soil mass, the depositional and
weathering history, climate (when the fissures were formed and now). It is useful to log, and classify the fissures
systematically.
McGown et al (1980) describe a method for describing soil fabric, including sedimentary layering, fissures, joints
and shears. Walker et al (1987) provide a fissure classification system which is simple and practical to use. The main
components are shown in Figure 38. Figures 39 and 40 show examples of fissured soils

Figure 38 : Fissure Classification Explanation Sheet (Walker et al 1987, after Coffey and Partners)

(a) View of side of test pit

(b) View of pieces of the fissured soil

Figure 39 : Weathered Tertiary clay, which exhibits shrink-swell (desiccation) induced fissures, the major
sets being inclined at about 20o to the horizontal, consistent with passive failure on swelling (a)
view of side of test pit, (b) view of pieces of the fissured soil.

Figure 40 : Fissures in pieces of excavated weathered volcanic tuff. The fissures are induced by desiccation
when the ground surface was exposed above sea level.
Fissures are caused by one or more of the following:
(a) Stress relief due to erosion or retreat of glaciation
(b) Freeze-thaw
(c) Shrink-swell (desiccation) due to seasonal moisture content changes
(d) Differential consolidation and settlement in sedimentary soils
It is important to recognise that the fissures are often formed in a different geological and climatic condition to the
present eg. in periods of glacial retreat, or when the sea level was lower than now. The fissured soils may in some
cases be overlain by more recent sediments which are not fissured (eg. Thorne, 1985).
The shear strength of fissured soils is dependent on

The peak, fully softened and residual strength of the soil substance

The continuity, orientation, shape and, spacing of the fissures and nature of the fissure surface.

These are in turn related to the origin of the fissures.


For example, smooth, planar fissures may only reduce the soil mass strength to the fully softened strength, but
polished and slickensided, planar or undulose, continuous fissures, will reduce the strength in the direction of the
fissures to near the residual strength.
It has been common practice to adopt the fully softened strength to approximate the mass strength of the soil. This
is based on back analyses of slides in English clays (Skempton 1977, Chandler 1984). The authors experience is that
in some soils, eg. fissured residual basalt and tuff (MacGregor et al 1990, Moon (1992) and fissured marine clays
(Thorne 1985), there was evidence from back analysis of failures, or from laboratory testing and mapping of the
fissures, that the mass strength could be considerably lower than the fully softened. Table 6 summarises the data. Raby
Bay and Plantes Hill both had failures which could be back-analysed (albeit with rather poor information on pore
pressures at the time of failure).
Stark and Eid (1997) present data from 14 cases which show back-analysed mobilised shear strengths for fissured
clays as being about the average of the fully softened and residual strengths for soils with a liquid limit between 50130%. They explain this in terms of progressive failure, rather than the nature of the fissures controlling the shear
strength. Potts et al (1997) explain the reduction of strength in slides in English clays to progressive failure. While
these may be reasonable explanation in some situations, eg. where the fissures are rough or smooth, and the soil

substance strain weakening, it is not sufficient where the fissures are polished or slickensided, and reasonably
continuous.

TABLE 6 : SHEAR STRENGTHS OF FISSURED SOILS FROM THREE PROJECTS IN AUSTRALIA


Project

References

Shear Strengths
Residual
Adopted

Peak

c'
Raby Bay(1)
Moon (1992)
5
Plantes Hill(2)
MacGregor et al (1990)
10
Thorne (1984)
25
Botany Bay(3)
(1) Residual soils derived from tuff. Strengths are those

'
c'R
24
2
25
0
24
0
used while the first

'R
c'
'
8
0
13
11
5
18
13
5
15
author was involved in project.

One failure was backanalysed to give c ' = kPa, ' = 13o.


(2) Residual soil derived from basalt.
(3) Marine clay fissured by desiccation and differential consolidation.
It is recommended that the following procedure be followed to assess the shear strength of fissured clays.
(a) Map the fissures for orientation, continuity and surface characteristics as detailed in Figure 38.
(b) Analyse the orientation using stereo projection methods as one does for rock joints (MacGregor et al (1990)
give an example).
(c) Carry out laboratory tests to determine the peak, fully softened and residual strengths, and if possible, the
strength along fissures (Thorne (1984) describes this procedure).
(d) Assess the strength of the soil mass parallel to the fissure surfaces accounting for the continuity and nature of
the fissure surfaces eg. polished slickensided fissures would be assigned residual strength, and if they were
50% continuous, the strength determined as the average of residual and fully softened strength. Normal to the
fissures, fully softened strengths would apply. For rough fissure surfaces, fully softened strength would apply.
This procedure is highly judgemental, and should where practicable, be backed up with back analyses of failures in
the same geological materials.
2.3.4
Cracking and Softening
Cracking of soil slopes can have a major influence on the stability:

By allowing water to enter the slope from the surface, increasing pore pressures generally, but more
particularly as pore pressures in the crack.

Allowing softening of the soil around the crack, reducing the undrained strength.

By the crack itself having zero strength if it stays open.


Piezometers installed in a slope will often not intercept the cracks, or be read too infrequently to measure the
transient pore pressures in the cracks.
It is important that these effects be properly modelled, particularly for small slopes, or for the upper parts of large
slopes, eg. embankment dams. It is also important that the limit equilibrium computer program which is used for the
analyses can model the cracks correctly. Simplified analysis methods, such as using zero strength but saturated unit
weight in the cracked zone, can give misleading factors of safety.
2.4
Failure Mechanics of Large Natural Landslides
2.4.1
The importance of large landslides
Large natural landslides (with volumes from say 1 million m3 to more than 1,000 million m3) have a particular
importance because they have the ability to cause damage and loss of life in single events eg. Vaiont, Italy 1963 caused
1,900 deaths, Mount Huascaran, Peru, 1962, 4,000 to 5,000 deaths, and in 1970, more than 18,000 deaths (Schuster
1996). There are also significant economic effects, including the direct cost of damage from the landslides, but more
commonly, the cost of investigation, monitoring and remedial works to reduce the risk, and sterilizing land for
development purposes.
This has been a particular concern to those who manage the reservoirs for hydro-electric, irrigation and water supply
dams, particularly since the landslide at Vaiont virtually filled the reservoir and caused a wave which overtopped the
dam to a depth of more than 100 meters (Riemer 1995).
2.4.2
The major issues and how they have been studied and addressed
The failure planes of large landslides in rock do not always have simple geometric shapes. Frequently, the
combination of discontinuities at the base of the landslide results in a polygonal shape. In consequence, the slide mass
undergoes substantial internal deformation before it can start moving and when it is moving. The internal deformation

can constitute a major proportion of the resisting force for first time and subsequent failures. Once the slide has moved,
a proportion of the internal resistance gets lost and the slide will reactivate more easily.
The process of internal deformation and strength deterioration has been observed in the Vaiont slide (Hendron and
Patton 1985). The surface of rupture follows folded strata. The upper part dips about 35o to 45o. With a short
transition, the bedding at the front of the slide becomes sub-horizontal. A seismic survey, conducted in 1959 indicated
extraordinarily firm in-situ rock (Mller 1964) with only 10 to 20m superficial, loose material. When lake filling
began in 1960, deformations measured on the ground surface of the slide accelerated, the perimetral crack opened and
shallow slides occurred at the front. A second seismic investigation, carried out in December 1960, determined the
thickness of the loose layer as 30 to 50m in the eastern part of the slide and 70 to 150m in the western part. The seismic
velocity in the compact rock at depth was twice as high as that in the deteriorated rock. A process of fracturing and
yielding in the upper layers of the slide mass had taken place over the time between the two surveys. Rotation and
toppling failure at the front of the slide mass above the gorge gave evidence of the shear deformation. During reservoir
filling, microseismic activity in the area of the landslide was recorded. For the final failure on 9 October 1963, a
seismic record exists. According to an interpretation of the seismogram, a phase of brittle fracture with a duration of 60
to 70 sec preceded the actual failure which took only 20 sec (Jaeger and Cook 1969).
In many large rock slides, the slide mass has partially converted into a blocky and frequently chaotic debris (for
instance the Cromwell Gorge landslides in New Zealand, the Alexis slide on the Polyphyton reservoir, Riemer (1995)
and the landslides at the Thissavros dam site in Greece). Heim (1932) and Mller & John, (1963), consider the
progressive development of a failure zone as a characteristic feature, of large landslides. The failure zone can develop
through breaking, rotating and grinding of rock blocks and produce a new material in the slope whose appearance and
properties are difficult to predict. With this effect, the strength of a rock slope becomes a function of strain and time.
The appearance of the debris affects the hydrogeological regime. In particular, surface infiltration in the debris is
enhanced. Secondary, superficial slides can develop in this debris, particularly near the toe of the debris.
Another common feature of large landslides is segmentation. The segmentation in some cases derives from relict
structure of the rock mass, in other cases it reflects a segmentation of the basal failure plane. When re-activated, these
segments may move independently and a long period of jostling and re-arrangement of the segments follows before
uniform conditions are re-established (eg. Jackson slide at the Cromwell Gorge in New Zealand, Alexis slide in
Greece).
The major concern in many cases is whether a slope which has been identified as an existing landslide, may become
fast moving (eg. when the reservoir for a dam is filled). Occasionally, a slope is identified as a potential slide (eg.
Dutchmans Ridge, Moore and Imrie 1992), and again the main question is whether it will become fast moving.
For individual projects the landslides or potential landslides have been investigated and analysed using the normal
geological and geotechnical methods. Judgements have been made by those involved based on the mechanics of
sliding, calculated factors of safety, and results of monitoring of movements, pore pressures, etc.. In many cases, where
the consequences of failure have been high, remedial works consisting of drainage (eg. into tunnels), and slope
modification works are carried out. Usually the increase in factor of safety is small (less than 10%), and slide
movement rates reduce, rather than stop (Imrie et al 1992, Jennings et al 1992). On-going monitoring, and maintenance
of the drainage systems are necessary.
2.4.3
Studies of the performance of large landslides
There are literally hundreds of papers which discuss individual large landslides eg. in the Proceedings of the
International Symposiums on Landslides. Few of these papers give much detail. There have been several noteable
studies to generalise the information eg. Heim (1932 (a remarkable book for its time), Eisbacher and Clague (1984),
Hs (1975), Voight (1978) and Schuster (2000). However these do not specifically address the issue of how to identify
those landslides which may become fast moving.
Studies to do this have been carried out by Hungr and Evans (2000) and J. Glastonbury (Glastonbury and Fell 2000)
is part way through a study. Hutchinson (1987) addresses the issue. These papers are summarised below. It should be
recognsied that such studies are inherently difficult:

The landslides are large, often in remote areas, and spread around the world

The studies concentrate more on landslides which have become fast moving, rather than those which remain
slow moving and on slopes which have failed, rather than remained stable, ie. the sample is not a reflection of
the population of slopes world wide.

Often there is little available data. Many investigators have made a few observations, rather than carrying out
detailed studies.

Those slides which have been studied in detail are rather few in number and the detailed data is sometimes not
available (for various reasons, including intellectual property and legal and commercial concerns).

The evidence has often been destroyed by the landslide (for those which became fast moving).

Relying on papers and reports, rather than first hand, intimate knowledge has potential pitfalls. This has been
pointed out by Hendron and Patton (1985) for the Vaiont landslide. Even the current authors have different
views on the mechanism of some slides.

2.4.4. A summary of views on ways to assess the likely velocity of movement of large landslides
(a) Hutchinson (1987) suggests the following mechanisms involving brittleness within the slide mass can lead to
high velocity landsliding

Internal shearing through brittle rock in compound (non circular or planar) slides as shown in figure 41.

Figure 41 : Typical internal shears required to permit


movement in a non-circular slide (Hutchinson 1987)
Hutchinson indicates this mechanism was present in the Vaiont slide.

Buckling, wedging and shear on low angle discontinuities in translational slides in jointed rock as shown in
Figure 42.

Reduction in the strength of pre-existing shears (either of tectonic or landslide origin) with further movement.
This may occur because the existing shear surface may have gained strength by cementation; weathering (eg.
hydrothermal alteration or other chemical weathering); or because the shear surface had not yet developed true
residual strength with asperities remaining or incomplete development of the shear surface.

Development of physical restraints due to internal displacements in the landslide, disrupting what were shear
surfaces at residual strength, and forcing further failure through intact, or less sheared rock.

Coalescence of landslides, where for example sliding of the lower part of a slope may remove support for the
upper part which had not sheared previously.

Figure 42 : Translational slides in jointed rock. Failure mechanicsm involving


(b) buckling (b) wedging (c) shear on low angle discontinuities (Hutchinson 1987)
(b) Hungr and Evans (2000), (also Hungr and Evans 1993, Thurber Engineering 1992) investigated originally 57
cases of sliding in rock slopes with a volume of 1 million m3, or deposit areas of at east 0.1 km2. The cases were mostly

taken from North America and Europe, but some others were included. They identified 9 classes or styles of failure as
shown in Table 7.
Weak and strong rock are not distinctly defined and the boundary between them is in any case not clear. More
details are given in the paper but can be summarised as:

Rock slumps occur in weak rocks such as shale, marl or tuff which are isotropic, or structured so the weak
direction is horizontal or dips gently opposite to the slope direction. They are usually ductile (slow).

Rock collapse (or rockfall using the classification in Table 7) occurs in steep (usually greater than 45o)
slopes, most often in volcanic, plutonic or low grade metamorphic rocks, but may take place in massive
sandstones or carbonates, they are catastrophic (defined by Hungr and Evans as landsliding involving
extremely rapid velocities (>5m/sec) and maximum displacements of more than 10% of the landslide source
area).

Translational rock block or wedge slides are often catastrophic but may be slow in weak rocks if the stress
levels are low, and the slopes not steep.

TABLE 7 : PROPOSED CLASSIFICATION OF FAILURE MECHANISMS OF LARGE ROCK


SLIDES.CASE STUDIES SHOWN IN ITALICS ARE DESCRIBED IN THE ORIGINAL REFERENCE
(HUNGR AND EVANS, 2000)
Structural
control
Dominant
mechanism
Constraint
Type

No systematic
structural
control

Systematic structural control


Sliding

Toppling

Unconstrained
Constrained
No internal
Internal
Laterally
At toe
At crest
Flexural
Block
deformation
deformation
Name
C
A
D
E
F
G
H
I
Block slide Compound
Block slide
Rock slump
Structurally
Half wedge
Flexural Block topple
with toe beak
Wedge slide
defined
slide
topple
B
Rock collapse
out
compound slide
Slow, rotational Slow (Gaillard Slow (Portuguese Slow (Butte,
Slow
Slow
Slow
Typical
Cut) Catastrophic
Bend)
Montana)
(Downie Slide) (Smoky
(La
movement
behaviour in
River,
Clapire)
(Massif de Plat) (Mt. Granier)
weak rock
Alberta)
Catastrophic
Catastrophic,
Catastrophic,
Catastrophic Catastrophic Catastrophic
Catastrophic
Typical
(Val Pola)
(Sherman (Frank Slide)
(Mystery
collapse, steep
limited prelarge pre-failure
behaviour in
Glacier)
Creek)
slopes, limited
failure
deformation
strong rock
(Vaiont)
volume (Elm)
deformation
(Goldau)

Flexural topples occur in anisotropic strength rocks such as schist, slate, phyllite or slate. They are
predominantly ductile, and Hungr and Evans indicate they know of no case leading to catastrophic failure.

Block toppling is usually catastrophic and occurs in steep slopes and strong, blocky rock masses. However
there are cases of large scale slope deformation; with secular creep (Chigra, 1992).

Each of the four remaining types: structurally defined compound slide (D) (of which Vaiont is an example),
half wedge (E), block slide with toe breakout (F) and compound slide with weak basal surface (G) exhibits a
range of behaviour, depending primarily on the strength of the rock mass and steepness of the original ground
surface. The type of behaviour (slow or catastrophic) is determined by a critical value of rock mass strength,
which is probably different for each of the failure types, and yet to be defined.
(c) J. Glastonbury is studying the pre and post failure deformation of large landslides, and constructed rockslopes.
The approach being taken is to assemble case studies where there is a reasonable level of detail of the geometry, and
geology, and in some cases, hydrogeology and displacements. The intention is to look in more detail at the rock
mechanics aspects of the cases than has been done in the other studies.
The work is still in progress, but the following points are presently apparent (Glastonbury and Fell 2000,
Glastonbury and Douglas 2000)).

Most rapid landslides are first time slides, rather than reactivated slides. The most common rapid failures are
translational slides, compound slides, rockfall and block toppling.

Rapid compound slides, such as Vaiont, require some brittle internal deformation of the slide mass.

Many rapid slides have involved brittle failure along lateral margins of the slide mass.

Translational slides are likely to be rapid if strain weakening on the rupture surface results in a residual friction
angle significantly less than the rupture surface inclination. They are likely to be slow (as evidenced by certain
geomorphic features) if the residual friction angle is equal to or slightly greater than rupture surface inclination.

Flexural toppling is often associated with the development of slow moving slides (often of general rotational
form).

Translational and compound slides which have a disaggregated slide mass (with blocks of rock separated by
broken material) are less likely to become rapid slides.
Many of the more obvious slides
(geomorphologically), and the landslides which are being monitored because of this in relation to reservoirs,
are likely to be disaggregated.
The seismic P wave velocity, would seem a good way of assessing the degree of disaggregation but it is not yet
clear what the limiting velocity would be. However slide material with a P wave velocity less than 1,600m/sec has the
seismic properties of soil, so would seem likely to represent totally disaggregated slide material (see Table 12). Clearly,
a more reliable way to estimate disaggregation is to develop a more detailed subsurface model by drilling, sampling,
tunnels etc.

Many of the well-documented rapid slides have occurred in areas which were subject to glaciation
approximately 10-20,000 years ago. In contrast, slow fluvial development of valleys is often associated with
slow moving slides.

Strongly anisotropic rocks (such as schists and phyllites) and strongly regionally folded, sheared and faulted
terrain rarely shows rapid landsliding.

Earthquakes M6 to M7 appear to have been the trigger for a number of rapid, first time slides. Small (M3 to
M4) earthquakes have occurred just prior to other slides, but these may be seismicity induced by pre-cursor
movement of the landslide.
There is some evidence that for disaggregated slides, including translational, earthquakes do not lead to rapid, large
displacement sliding.

Even modest slide velocities, (tens of metres per day), can cause major problems with rivers being blocked eg.
Thistle, Paonia slides. More work is needed to determine why these mid-level velocities occur.

Most, if not all landslides which become rapid, show some signs of movement prior to the main event. Heim
(1932) and Leroueil (2000) gives a number of examples, and others eg. Evans, (personal communication),
think this is the case. The exception could be first time slides induced by large earthquakes. The critical issue
will be recognising the warning signs. This is being studied further.

3.

GEOTECHNICAL INVESTIGATION REQUIREMENTS AND METHODS

3.1
Investigations Required for Different Classes of Landslide, and Level of Investigation
3.1.1
Staging of Site Investigations An Iterative Approach
Geotechnical investigations are best carried out in stages, and with allowance for iteration. Figure 43 shows the
activity flow in site investigation.

Figure 43 Activity flow in site investigations (Fell et al 1992)

The following notes relate to the seven activities shown in Figure 43.
Activity 1. First the objectives of the work, or questions to be answered, are defined. These include both geological
and engineering questions.
Activity 2. Existing, or readily available, regional and local geological and other data relevant to the site are
collected and completed to give a tentative geotechnical model. Tentative answers to the questions asked in activity 1
are obtained where possible either from local knowledge or from rapid analysis, or both. New geological questions are
usually added at this stage, arising from the understanding obtained from the existing data.
Activities 3 and 4. To answer the remaining questions and to confirm the tentative answers, the various subobjectives and activities of Activity 5 to 7 are defined. The activities are related to time and money, respectively, in
activity charts and cost estimates, and set out in a report, seeking approval to proceed.
Activity 5. This relates to achieving all of the essentially geological sub-objectives defined in Activity 3. The
engineering geological model here implies a sufficient understanding of the regional geology, geological history, and
detailed site geology. The model is described in geotechnical language.
Activity 6. Field or laboratory tests, or both, are carried out to obtain quantitative values usually for engineering
properties of critical portions of the detailed site model. The test results are assessed in the light of the geological
model and from this, certain values or ranges of values may be adopted as realistic, and included in the model which is
now termed the geotechnical model.
Activity 7. Engineering analyses are then carried out involving the proposed structure and the site geotechnical
model. For the parts of the model without test results, assumptions are made based on precedents and knowledge of the
likely variability of the types of materials present. It is desirable that the analyses provide answers in terms of
probability (likelihood) of failure, as well as factor of safety.
Usually many questions will have to be answered by judgement, based on experience in similar geological
situations. If at the end of Activity 7, all questions are answered with sufficient confidence, the investigation is
complete. If not, further cycles of investigation are carried out until the required level of confidence is reached. The
law of diminishing returns is applied. One of the most difficult tasks may be to decide if a landslide hazard exists
which needs to be investigated. This question was strongly debated in relation to the Polyphyton and Thissavros
reservoirs in Greece. In both cases, international experts had rejected the hypothesis of the existence of large landslides.
In Polyphyton some minimum precautions were nevertheless taken and the consequences for the project were minor.
At Thissavros, very substantial and time consuming stabilization was required which perhaps would have been avoided
if the project design had already taken the landslides into consideration. This was a similar issue with the Downie Slide
in Canada.
The incidence of landsliding, and the mechanics of landslides are often closely related to the underlying geology and
topography (Fell et al 1992, Patton and Hendron 1974, Hunt 1984). Because of this, and the need to get the local
geology and hydrogeology in the overall context, it is important that geological and geotechnical investigations proceed
from the regional to local, and then site specific studies using a multidisciplinary team.
The regional studies will usually include some or all of the following:

Assembly of available topographic maps

Examination of regional geology maps, cross sections and reports

Interpretation of air photography (and satellite imagery if applicable)

Ground reconnaissance over previously mapped areas, and possibly remapping of some important areas.

Assembly of the historic record of landsliding.


The local and site specific studies are likely to include some or all of:

Detailed geotechnical mapping

Definition of the stratigraphy and structure/fabric

Geophysical methods (where applicable)

Direct investigation of subsurface conditions including trenching, test pitting, drilling, aditing sampling, in-situ
testing

Downhole logging and inspection where applicable

Laboratory testing

Plotting of all data, development of a geotechnical model

Reporting.
Geotechnical landslide investigation sometimes has to cover large areas and huge masses of rock and soil (on the
order of km3 in the case of Downie slide in Canada and 9 Mile Creek slide in New Zealand). This condition reinforces
the need for an iterative approach with progressive refinement to the program of geotechnical investigation. Rapid
surveys on large scale (geological, geo-morphological, geophysical) should initiate the work and prepare for detailed
investigations (drilling, aditing). The reliable definition of the slide model, suitable for a correct geotechnical analysis,
depends on the size and the complexity of the landslide and may require a large volume of investigation. The
investigation program has to be well planned and managed to be cost efficient and to be successfully completed in good
time. To justify the respective investment of time and funds, a hazard and risk rating of existing and potential
landslides should be made, based on the findings of the basic surveys. In this sense, the landslides along Cromwell

Gorge (New Zealand) were divided into two groups, the potentially catastrophic slides and the minor slides. The
potentially catastrophic slides had to be studied in detail. When the work had advanced, a further subdivision of the
potentially catastrophic slides into first and second priority became possible. Eventually, after treatment, the
classification changed to distinguish, for the purpose of performance monitoring and risk management, into active
slides, creeping slides and dormant slides. Progressive revision of the geotechnical concepts and of the risk rating of the
investigated slide should be made to achieve technical and economical optimization of the investigation program.
3.1.2
Questions to be Answered and the Site Investigations Methods Available to Answer Them.
All site investigations, whether for natural or constructed slopes, should be carried out with a clear objective, and
with a set of questions to be answered. Stapledon (1995), and Sowers and Royster (1978) (in Turner and
McGuffey,1996) give examples of the questions to be answered.
Table 8 is developed from them, and is applicable to all classes of slopes, although clearly some questions are more
applicable to large, natural landslides and other larger slopes than to a constructed fill for example.

TABLE 8 : QUESTIONS TO BE ADDRESSED IN SLOPE STABILITY


AND LANDSLIDE INVESTIGATIONS
1. TOPOGRAPHY?
2. GEOLOGICAL SETTING?

3. HYDROGEOLOGY?

4. HISTORY OF MOVEMENT?

1.1
1.2
2.1
2.2
2.3
3.1
3.2
3.3

3.4
3.5
3.6
4.1
4.2
4.3

4.4
5. GEOTECHNICAL
CHARACTERISATION OF THE
SLIDE OR POTENTIAL SLIDE?

5.1
5.2
5.3

6. MECHANISMS
AND
DIMENSIONS OF THE SLIDE OR
POTENTIAL SLIDE?

8. ASSESSMENT OF STABILITY?

6.1
6.2
6.3
6.4
7.1
7.2
7.3
7.4
7.5
7.6
8.1

9. ASSESSMENT OF
DEFORMATIONS AND TRAVEL
DISTANCE?

8.2
9.1
9.2
9.3

7. MECHANICS OF SHEARING
AND STRENGTH OF THE
RUPTURE SURFACE?

In the landslide source and potential travel path


Effect and timing of natural and human activity on the topography
Regional stratigraphy, structure, history (eg. glaciation, sea level
submergence and emergence)
Local stratigraphy, slope processes, structure, history
Geomorphology of slope and adjacent areas
Regional and local groundwater model?
Piezometric pressures within and around the slide?
Relationship of piezometric pressures to rainfall, snowfall and
snowmelt, temperature, streamflows, reservoir levels, both seasonally
and annually?
Effect of natural or human activity?
Groundwater chemistry and sources
Annual exceedance probability (AEP) of groundwater pressures
Velocity, total displacement, and vectors of surface movement?
Any current movements and relation to hydrogeology and other natural
or human activity?
Evidence of historic movement and incidence of sliding eg. lacustrine
deposits formed behind a landslide dam, shallow natural slides, or
failures of cuts and fills
Geomorphic or historic evidence of movement of slope or adjacent
slopes
Stage of movement (pre failure, post failure, reactivated, active)
Classification of movement (eg. slide, flow)
Materials factors (classification, fabric, volume change, degree of
saturation)
Configuration of basal, other bounding, and internal rupture surfaces?
Is the slide part of an existing or larger slide?
Slide dimensions, volume?
Is a slide mechanism feasible?
Relationship to stratigraphy, fabric, pre existing rupture surfaces
Drained or undrained shear?
First time or reactivated shear?
Contractant or dilatant?
Saturated or partially saturated?
Strength pre and post failure, and stress-strain characteristics.
Current, and likely factors of safety allowing for hydrological, seismic
and human influences?
AEP of failure (factor of safety  
Likely pre failure deformations?
Post failure travel distance and velocity?
Likelihood of rapid sliding?

Table 9 lists the site investigation methods which are available and their application to answering the questions
posed in Table 8 Also listed are references which give good overviews or examples of application of the methods.

TABLE 9 : SITE INVESTIGATION METHODS


Method

Description

Topographic
mapping and survey
Regional Geology

Regional and project area


contour plans at appropriate
scales
Existing
geology
maps,
regional mapping for the
project

Assists in Answering
Questions
1.1, 1.2,

2.1, 3.1, 4.4

Geological mapping
of project area

Mapping
stratigraphy,
structure, faults and shears,
weathering,
groundwater
features, soil fabric

2.2, 3.2,3.3, 3.5, 4.3,


6.1, 6.2, 6.3, 6.4, 7.1,
7.3, 8.1, 9.1, 9.2, 9.3

Geomorphological
mapping

Mapping topographic, surface


shape, drainage, and landslide
features such as scarps,
hummocky ground, deflection
of rivers, presence of rapids.
Inspection of satellite imagery
to define regional geological
features, faults; hydrogeology
vegetation, deformation (by
interferometry)
Aid to geomorphological and
geological mapping, terrain
classification
Assessing the frequency and
nature of past movements
from written and verbal
history
Dating landslide movements
by
radiocarbon
dating,
stratigraphic relationships and
tree ring dating
Use of surface seismic P and S
wave (surface, crosshole,
surface
to
downhole),
electrical
resistivity:
downhole
induction,
resistivity,
self potential,
gamma, neutron
Backhole,
excavator
or
bulldozer trenches, and pits
for investigation and sampling
subsurface conditions
Drilling and sampling using
augers, washboring, coring,
etc., and sampling with thin
walled samples, core barrels.
Etc., oriented core samples

1.2, 2.3, 3.1, 4.1, 4.2,


4.3, 4.4, 5.1, 5.2, 5.3,
6.1, 6.2, 6.3, 8.1, 8.2,
9.1, 9.2, 9.3

Satellite
imagery
interpretation

Air
photograph
interpretation
Historic record

Dating
movements

past

Geophysical methods

surface
and
downhole

Trenches
Pits

Drilling/boring

2.1

2.1, 2.2,
4.4, 5.1,
6.3, 8.1
2.2, 4.1,
5.2, 8.1,
9.3

2.3, 4.1, 4.3,


5.2, 6.1, 6.2,
4.2, 4.3, 5.1,
8.2, 9.1, 9.2,

References
Walker et al (1985), Fell et al
(1992)
Fell et al (1992), Keaton and De
Graf (1996), Geological Society
Engineering Group (1972, 1982),
Stapledon (1995), Hutchinson
(1995), Galster (1992), Dearman
and Fookes (1994)
Fell et al (1992), Keaton and De
Graf (1996), Geological Society
Engineering Group (1972, 1982),
Stapledon (1995), Hutchinson
(1995), Galster (1992), Dearman
and Fookes (1994)
Geological Society Engineers
Group (1972), Dikau et al (1996),
Brunsden et al (1975), Rib and
Laing (1978), Hutchinson et al
(1991),
Soeters and Van Westen (1990)

Rib and Laing (1978), Hunt


(1984), Soeters and van Westen
(1996), Lillesand (1987)
Turner and Jayaprakash (1996)

2.2, 4.1, 4.2, 4.3, 5.1,


5.2, 8.1, 8.2, 9.1, 9.2,
9.3

Stapledon (1995), Worsley (1981)

2.2, 3.1, 5.3, 6.1, 6.2,


6.3, 3.5

Fell et al (1992), McGuffey et al


(1996), Stapledon (1995), Hanna
(1985), Clayton et al (1995)

2.2, 3.2, 3.5, 4.3, 5.3,


6.1, 6.3, 6.4, 7.1, 7.3,
7.5

Walker et al (1985), Fell et al


(1992), Stapledon (1995)

2.2, 3.1, 3.2, 5.3, 6.1,


6.3, 7.1, 7.3, 7.5

McGuffey et al (1996), Hunt


(1984), Fell et al (19920, Walker
et al (1985), Stapledon (1995)

TABLE 9 SITE INVESTIGATION METHODS (CONT.)


Method

Description

Downhole inspection

Shafts, tunnels and


adits
In-situ testing of
strength
and
permeability

Monitoring
pore
pressures,
rainfall,
snowfall, snowmelt,
temperature
and
seepage flows

Monitoring
displacements
deformations

of
and

Laboratory
testing
relating to shear
strength

Back-analysis
of
stability to determine
shear strength

Inspection/testing of the sides


of boreholes using downhole
cameras, video, borehole
impression, ultrasonic logging
Excavation of shafts or
tunnels (adits) for mapping
and sampling
(a) Standard Penetration Test
(SPT), Cone Penetration
Test (CPT), Piezocone
(CPTU)
(b) Vane Shear (Soft clays
only)
(c) Borehole
permeability,
pumping and packer tests
Installation of piezometers to
monitor
pore
pressures/groundwater
pressures, and relation to
rainfall, snowfall, snowmelt
and temperature
Install
seepage measuring weirs.
Measurement of surface and
subsurface displacements and
deformations
by
surface
survey,
borehole
inclinometers, extensometers,
tiltmeters,
crackmeters,
acoustic monitoring
Laboratory triaxial, direct
shear, ring shear, direct simple
shear to determine peak and
residual
strength;
water
content, density, classification,
clay
content
and
clay
mineralogy
Back-analysis of moving
landslide to assess shear
strength

Assists in Answering
Questions
3.1, 5.3, 6.1, 7.1

2.2, 3.1, 3.2, 3.3, 3.5, 4.1,


4.2, 4.3, 5.1, 5.2, 5.3, 6.1,
6.2, 6.3, 6.4, 7.1, 7.3, 7.4
2.2, 5.3, 6.1, 6.3, 7.1, 7.2,
7.4, 7.6, 8.1

References
Clayton et al (1995)

Gillon et al (1992), Stapledon


(1995)
DeCourt et al (1988), Drumright
et al (1996), De Beer et al
(1988), Lunne et al (1997)

7.2, 7.6, 8.1


2.2, 3.3, 3.5, 8.1
Fell et al (1992)
3.1, 3.2, 3.3, 3.4, 3.5, 3.6,
5.3, 7.5, 8.1, 8.2

Dunnicliff (1995), Walker and


Mohen
(1985),
Stapledon
(1995),
Mikkelsen
(1996),
Bromhead (1986), Fell et al
(1992), MacFarlane et al (1992)

4.1, 4.2, 5.1, 5.2, 6.1, 6.2,


6.3, 7.1, 8.1, 8.2, 9.1, 9.2,
9.3

Mikkelsen (1996), Dunnicliff


(1995), Kovari (1990)

5.3, 6.1, 6.4, 7.2, 7.3, 7.4,


7.5, 7.6, 8.1, 8.2, 9.1, 9.2,
9.3

Head (1980, 81, 85), Fell and


Jeffery (1985), Fell et al (1992)

7.2, 7.6, 8.1

Not all methods are applicable to all classes of slopes, or to all stages of investigation. Table 10 lists the applicability of
the methods to typical classes of slope problem. It will be seen that for:

Shallow natural landslides there is a reliance on geological, topographical geomorphological and historic
information without detailed drilling, sampling, laboratory testing and analysis.

Medium natural landslides have a similar emphasis on geology, geomorphology and historic records, but a
greater emphasis on sub-surface exploration, sampling, monitoring of pore pressures and laboratory testing.

Large natural landslides there is a greater emphasis on monitoring of deformations and groundwater pressures,
less emphasis on laboratory testing and more on back-analysis to assess strengths.

Existing cuts and fills, there is a reliance on topography, geology, geomorphology and historic performance.
For larger structures, more subsurface investigation, sampling and laboratory testing will be carried out.

New cuts and fills, there is a reliance on the performance of structures in similar conditions, and monitoring
post construction. It is essential that geomorphological studies be included to identify existing natural
landsliding.

Embankments and cuts in soft clays, the emphasis is on stratigraphy and strength, often obtained by in-situ
tests, and careful drilling, sampling and laboratory tests.

TABLE 10 : APPLICATION OF SITE INVESTIGATION METHODS TO SLOPE CLASSES


Natural Slopes
Constructed Slopes
Site Investigation
Method
Small/Shallow Medium Large Existing Cut Existing Fill New Cut New Fill Soft Clay
Topographic mapping
A
A
A
A
A
A
A
A
and survey
Regional Geology
A
A
A
A
A
A
A
A
Geological mapping of
B
B
A
A
B
A
B
C
project area
Geomorphological
A
A
A
B
B
B
B
D
mapping
Satellite
imagery
D
D
C
D
D
D
D
D
interpretation
Air
photograph
A
B
A
C
C
C
C
C
interpretation
Historic record
A
B
B
A
B
B(2)
B(2)
B(2)
Dating past movements
B
C
B
D
D
D
D
D
Geophysical methods
C
C
B
C
C
C
D
C
Trenches and Pits
B
A
B
B
B
B
B
C
Drilling/boring
C
A
A
C
B
B
B
A
Downhole inspection
C
B
B
C
D
C
D
D
Shafts and tunnels
D
C
B
D
D
D
D
D
In-situ testing of strength
C(3)
C(3) C(4)
D
B(3)
C
C
A(3)
and permeability
Strength
and
permeability monitoring
C
A
A
A
A
C
C
A(5)
pore pressures, rainfall
etc.
Monitoring
of
C
B
A
B
B
B(5)
C(5)
A(5)
displacements
Laboratory testing
C
A
B
B
B
B
C
A
Back analysis of stability
C
B
A
C
B
B(2)
C(2)
C(2)
Notes: (1)
A Strongly applicable,
B Applicable,
C May be applicable,
D Seldom applicable
(2)
In similar areas
(3)
SPT, CPT, CPTU
(4)
Permeability
(5)
During construction
3.2

Comments on Some Investigation Methods


The following are some detailed comments on some of the site investigation methods.

3.2.1
Geomorphological Mapping
Geomorphological mapping is an essential part of most site investigations for natural and constructed slopes. For
the latter, it is to assist in defining the overall geotechnical environment in which the slope is or is to be constructed, and
in particular to assist in identifying the presence of existing instability at the site or nearby.
Geomorphological mapping does not have to be expensive, or done by sophisticated methods. Valuable maps can
be made using simple hand inclinometers, hip chain and sketch maps. The system of symbols does not have to be
complicated. Figure 44 shows an example.
The mapping is a valuable base for more detailed subsurface investigations.
3.2.2
Historic Record
The historic record of landsliding in an area or of a particular slide can be the best guide to future performance. This
can take several forms:

Published information in technical literature, history of local communities, newspaper articles.

Records of instability along roads and railway lines.

Data obtained from successive photography.

Data inferred from geomorphological, stratigraphic and carbon dating.

Figure 44 Large scale geomorphological plan of a landslide area


at OShannassy dam, near Melbourne.
(Fell et al, 1992, Courtesy of Jeffery and Katauskas and Melbourne Water).
Often, the historic record can be used, along with an assessment of the climatic or other triggering conditions, as a
guide to the frequency of landsliding, and used as a guide to the future likelihood of sliding, in a probabilistic
framework. This is valid provided the conditions controlling landsliding have not changed.
The historic record can also be used as a guide to likely velocity and travel distance of movement, and is often more
reliable than the available methods of calculation. For example, if the historic record shows that road fills in a certain
geological and climatic environment constructed without compaction occasionally become debris flows on failure (due
to collapse on shearing) it may be wise to assume that similar fills may also experience such flows, rather than to
attempt to quantify the mechanics by laboratory testing and analysis. Such an approach has been taken in regard to fills
constructed prior to 1980 in Hong Kong, and the fills are systematically removed, or the outer parts reconstructed so
they will not flow even if they fail (Morgenstern 1978, Wong et al 1998).
3.3.3
Surface Geophysical and Borehole Logging
(a) Surface, and surface to borehole methods.
Geophysical techniques have a great attraction for some because the methods are non-invasive, appear likely to
sample the mass as a whole and give good coverage at low cost, rather than in a point (as does sampling in a borehole)
and are high tech.
Unfortunately the reality is that there are few landsliding situations where the techniques are a great deal of value.
Stapledon (1995), McGuffey et al (1996), and Clayton et al (1995) give general assessments of the methods.
Table 11 is adapted from McGuffey et al (1996) to include the authors experience.

TABLE 11 : APPLICATION OF SURFACE BASED GEOPHYSICAL METHODS


(ADAPTED FROM MCGUFFEY ET AL 1996)
TYPE OF SURVEY
Electrical and electromagnetic
Electrical resistivity

Electromagnetic conductivity
profiling

Seismic
Seismic p wave refraction
profiling

Direct seismic (uphole,


downhole, and crosshole
surveys) p and s waves

APPLICATIONS

LIMITATIONS

Locates boundaries between clean


granular and clay strata, groundwater
table, and soil-rock interface

Difficult to interpret and subject to


correctness of the hypothesized
subsurface conditions; does not provide
engineering strength properties.

Locates boundaries between clean


granular and clay strata, groundwater
table, and rock-mass quality; offers
even more rapid reconnaissance than
electrical resistivity

Difficult to interpret and subject to


correctness of hypothesized subsurface
conditions; does not provide
engineering strength properties. Not
recommended for most landslide
investigation.

Determines depths to strata and their


characteristic p wave seismic
velocities

May be unreliable unless strata are


thicker than a minimum thickness,
velocities increase with depth, and
boundaries are regular. Information is
indirect and represents average values

Obtains p or s wave velocities for


particular strata, their dynamic
properties (s wave), and indirectly,
rock-mass quality

Data are indirect and represent


averages may be affected by mass
characteristics. Methods of
interpretation using tomography are
improving.
Use of expensive and sensitive
instruments in rugged terrain typical of
many landslides may be impractical;
requires precise leveling and elevation
data; results must be corrected for local
topographic features; requires detailed
information on topography and
material variations; not recommended
for most landslides investigations
Has limited penetration in clay
materials, and latextes. Not
recommended for most landslide
investigations.

Microgravity

Uses extremely precise measurement;


may locate small volumes of lowdensity materials utilizing very
sensitive instruments

Ground-penetrating radar

Provides a subsurface profile, locates


buried objects (such as utility lines),
boulders, and soil-bedrock interface

In the authors experience, the most useful technique is seismic refraction for large landslides. The method can
define the base of sliding if the undisturbed rock has a sufficient velocity contrast to the sliding material, and the
velocity of the slide mass can give a guide to the amount of disaggregation, which as discussed in Section 2.4, appears
to give a guide to the likelihood of rapid sliding. Table 12 presents the seismic refraction P wave velocities from the
Downie and Cromwell gorge landlsides.
Mller (1964) noted that seismic refraction velocities reduced as movement progressed in the Vaiont landslide, but
it did not indicate the possibility of a deep failure plane. However interpretation methods have improved since that
time. Seismic P and S wave tomography (cross hole or surface to hole) is another area which has advanced and
should have good application to larger landslides.
Geophysics should never be relied upon alone to determine subsurface conditions, and should always be supported with
drilling or tunnelling.
(b) Borehole logging
Borehole logging techniques are summarised in Mcguffey et al (1996), and Clayton et al (1995). Table 13,
summarises the usefulness of the main methods.
The limitations from a landslide investigation viewpoint are:

Many methods require an open hole - and often landslide masses are badly disturbed so the holes will not stay
open.

The sensitivity of the methods (to locate thin layers) is limited, so clay-rich zones on the rupture surface, for
example may be too thin to be located by some methods.

TABLE 12 : SUMMARY OF THE SEISMIC REFRACTION P WAVE VELOCITIES, DOWNIE


AND CROMWELL GORGE LANDSLIDES (STAPLEDON 1995, BASED ON DATA FROM
PITEAU ET AL 1978, BRYANT ET AL 1992 AND MACFARLANE ET AL 1992)
Zone
A

DOWNIE SLIDE
Description
Vp km/s
Rock-soil
0.4 0.8
mixture, dry,
local large voids

Moderately to
highly fractured
rock

Zone
Chaotic debris (dry)

CROMWELL GORGE SLIDES


Description
Vp
km/s
0.6 0.8
Disoriented
blocks in matrix
of sheared,
crushed and gouge
materials

1.25 2.8
Displaced rock

Large intact
blocks touching
or separated by
infill or sheared
or crushed
material

1.2 3.0

--------------------------------------Assumed base of slide movements----------------------------D

Undisturbed
bedrock

4.0 6.2

2.8 4.6

Insitu
rock

TABLE 13 : BOREHOLE LOGGING METHODS (ADAPTED FROM MCGUFFEY ET AL 1996)


Method
Ultra-sonic log

Parameter Measured

Thermal profile

Determination of sonic velocities,


attenuation of shear velocity in
shear zones, or imaging the sides
of the borehole wall
Temperature

Caliper log

Borehole diameter

Induction log

Electrical conductivity

Resistivity

Electrical resistivity

Self potential

Electrical potential from mineral


reaction and ground water flow

Natural Gamma

Natural Gamma radiation

Applications

Limitations

Shows fractures, other


discontinuities

Requires uncased hole.


Image less clear than
borehole camera.

Zones of inflow
(lower/higher
temperature water)
Infers lithology by hole
enlargement by erosion
during drilling by drill
fluid
Infer lithology, by clay
content, permeability,
degree of fracturing

Open hole not necessarily


required.

Infer lithology,
(particularly sand vs
clay)
Infers lithology,
oxidation/reduction
zones, subsurface flows
Infers presence of clay
and shale

Requires an uncased hole

Lower resolution than


resistivity log, can evaluate
unsaturated zone and pvc
cased boreholes.
Applicable only in
saturated zone; requires
open hole.
Applicable only in
saturated zone in uncased
borehole. Data difficult to
interpret.
Mud coating on borehole
wall may affect results.

TABLE 13 : BOREHOLE LOGGING METHODS


(ADAPTED FROM MCGUFFEY ET AL 1996) (CONT.)
Method

Parameter Measured

Gamma Gamma

Natural density

Neutron
Neutron

Moisture content (above water


table). Porosity (below water
table)

Borehole Camera or
Video

Visual image of fractures and


structure

Borehole impression
packers

Impression of fractures on a thin


film

Applications
Provides log of density,
from which lithology
may be inferred
Log of mositure content,
from which lithology, or
wetted zones may be
inferred.
Can identify structure.
May be able to interpret
dip of fractures.
Orientation of fractures

Limitations
Provides only density.
Health and safety issues.
Provides only water
contents. Health and
safety issues.
Requires uncased hole.
Images affected by water
quality, smearing of sides
of hole.
Requires open fractures,
slow to do.

3.3.4
Trenches, Pits, Shafts, Adits and Tunnels
There is no question that trenches, pits, and for large landslides, adits (tunnels) and possibly shafts, are an essential
part of landslide investigation. On smaller projects, trenches and pits are a cost effective way of exposing the strata,
allowing identification of colluvial material (as compared to weathered in-situ), exposure of soil fabric and rock
structure and the surface of rupture for mapping and sampling. Even the highest quality drilling and sampling is not as
good as the larger surfaces exposed in trenches, pits, shafts and adits.
In large landslides, the possibility of excavating adits or shafts in the framework of the investigation program should
be considered. In addition to the advantage of direct inspection and sampling of the rock and failure planes, adits and
shafts allow in-situ testing, can offer safe working sites for exploratory and drainage drilling and permit easy
monitoring of deformation and hydrogeologic processes. Functioning also as drains, the adits have in many cases
served the dual purpose of investigation and stabilisation. Although frequently ruled out because of the initial
construction cost, adits have repeatedly proven a particularly cost efficient method of investigation for large landslides.
Examples of the use of shafts and adits are given in Lewis and Moore (1988), Gillon et al 1992, MacGregor et al
(1990), Hutchinson et al (1973).
3.3.5
Drilling
Drilling of boreholes into landslides can be difficult because of the disturbed ground conditions which lead to caving
of the hole. Care should be taken in using drilling fluid so that there are not excessive losses which may lead to
increased piezometric pressures, softening or hydraulic fracture. Drilling into the core of embankment dams should
always use dry drilling techniques (eg. hollow flight auger, auger, dry cable tool) because of the risk of hydraulic
fracture (which may initiate internal erosion and piping) if wash boring, or other methods using water or air as a
circulatory medium are used.
The zones of crushed rock, gouge or clay seams are particularly important to be detected and sampled, but core
losses can hardly be avoided. In such cases, the recording of drilling parameters (eg. the drill progression rate, drill
fluid pressure) can assist in the interpretation of subsurface conditions. In any case, the borehole water levels should be
recorded regularly during sinking of the hole to check for the existence of perched or confined groundwater which will
be difficult to detect once the borehole is completed and the water level will display a mixed potential. MacFarlane et
al (1992) describe the use of depth of Hole Vs Depth of drill fluid to locate aquifers, assess whether the aquifer is
confined, and to assist in positioning piezometers.
Impression packers can be used to determine the orientation of specific features in the borehole. Oriented drill cores
will provide more comprehensive information but may be difficult to take from deteriorated rock. Alternatively,
optical, television or sonic logging can be carried out to measure the discontinuities.
Groundwater monitoring should normally be provided in completed boreholes.
3.3.6
Monitoring of Pore Pressures, Displacements and Deformations
The importance of good quality, well designed monitoring of pore pressures, displacements and deformations cannot
be over-emphasised. This applies to small as well as large slope stability studies. Instrumentation with appropriate
accuracy and response time have to be selected.
Some important points are:

Consider the detailed hydrogeological conditions before installing piezometers (or add more piezometers later
if monitoring shows this to be necessary). Relative permeabilities, both between strata, and the ratio of

SN O W D E PTH
(equivalent m m of w ater)

1000

500

0
S N O W D EP TH - W AT ER E Q U IVALE N TS

D R AIN A G E G A LLE R Y FLO W


(litres/m inute)

5000

4000

3000

2000

1000

0
D R A IN A G E G A LLE R Y FLO W
1720

P IEZ O M E TR IC LE VE L
(m etres)

horizontal to vertical permeability are critical to perching and compartmentalisation of ground water.
Examples of this are given in Bromhead (1986), Walker and Mohen (1985), MacFarlane et al (1992), Fell et al
(1992).
Take readings at sufficiently close time intervals, so the effects of seasonal rainfall, and in smaller slides,
rainfall intensity and duration during individual events can be monitored. As can be seen in the plots in
Figures 45 and 46, annual, or even monthly readings would have been quite misleading. In the Grand Maison
landslide shown in Figure 45, snowmelt occurs over a period of a month or so, resulting in tens of metres rise
in the piezometric levels even after drainage galleries have been constructed. In Figure 46, displacement
readings are shown to be strongly seasonal, so annual readings, as are often taken for large landslides, do not
convey the correct picture of rates of movement. Brand (1995) (Figure 47) gives an example of piezometric
pressures rising 6 metres in 4 hours in a slope in Hong Kong.

1710

1700

1690
1995

1996

1997

P IEZ O M E TR IC LE VE L

Figure 45 : Piezometric pressures recorded in the Grand Maison landslide


(Adapted from data supplied by Electricit De France)

Figure 46 : Time displacements and cumulative rainfall for the Divinei Rockslide
(Baumer and Schindler 1992)

Figure 47 : Example of a rapid piezometric pressure change in a Hong Kong slope,


probably typical of high permeability slopes elsewhere (Brand 1995)

The use of continuous recording, or at the simpler level, devices for recording the maximum recorded
piezometric level (eg. Halcrow bucket) is essential to get the complete picture.
Use piezometers, sealed in the borehole with carefully designed grout, so they only read pore pressures in the
length of hole they are designed for. Do not use wells holes back filled completely with sand or gravel,
since these will record the phreatic surface, and not allow for non-hydrostatic conditons.
Measure vectors of displacement, not just settlement. The vectors often reflect the shape of the rupture surface,
so can be a valuable guide to mechanism. Figure 48 shows how the surface displacements closely parallel the
rupture surface in a landslide on a railway line.

Figure 48 : Section through landsliding on a railway line.


Courtesy of LongMac Associates and State Rail Authority of New South Wales
4.

ANALYSIS OF PORE PRESSURES, STABILITY, AND PRE-FAILURE STRESSES AND


DEFORMATIONS

4.1
Analysis Required for Different Classes of Landslide and Level of Investigation
4.1.1
Some general points
A critical component of all analyses of seepage pore pressures, stability, and stresses and deformations in a slope,
whether natural or constructed is a good hydrogeological and geotechnical model.
For natural slopes, and slopes built on or into natural slopes, it can be very difficult to model the piezometric
conditions with any degree of accuracy, because the slopes are very complex, with preferential flow paths for water eg.
in open jointed rock, or in sandy beds in soil slopes; perched water tables; and the effects of construction and rainfall or
snowmelt. For these slopes, we must understand the geological environment and what it is likely to lead to
hydrogeologically and plan the site investigations to develop a detailed understanding. This is discussed in Stapledon
(1995), Fell (1995), Fell et al (1992), Walker and Mohen (1985), Patton and Hendron (1972).
For constructed fills, including dams, the situation is simpler, but still it is not easy to predict the pore pressures with
any degree of confidence because of the effects of compaction (giving potentially high - but difficult to predict) ratios of
horizontal to vertical permeability (Fell et al (1992), Bromhead (1986), the effects of partial saturation, and dependence
of permeability on the confining stress. Vaughan (1994), Bromhead (1986) point out that the permeability can vary by
1 or 2 orders of magnitude with confining stress and this can have a major effect on pore pressure distributions.
Because of this it is often necessary to develop a hydrogeological model, and instrument the slopes with piezometers
and monitor them over a sufficiently long period to establish, for example, the relationships between pore pressures and
rainfall or snowmelt. As discussed in Section 3.2, it is essential these instruments be read often enough to detect the
changes in pore pressure which occur in the slope. Even when slopes are well instrumented, care must be taken in
interpreting the data to allow for flow net effects (Iversen (1992), Bromhead (1986), Reid (1997) and
compartmentalisation of groundwater (eg. Macfarlane et al (1992), Walker and Mohen (1984). Alternatively, or in
addition, it may be necessary to engineer the pore pressures to make them more predictable by for example installing
sub-horizontal borehole drains into a cut slope, or constructing a drainage layer under a fill to intercept seepage from
the foundation before it enters the slope.
4.1.2
Methods of estimating pore pressures and analysis of stability for different classes of slopes.
The following summarises what are appropriate methods for estimating pore pressures and stability analysis of
different classes of natural and constructed slopes.
(a) Natural slopes shallow/small slides.
For slides less than say 5 metres thick, the factor of safety which can be calculated from the assumed geometry pore
pressures and shear strength, are very sensitive to the input parameters, and many would consider the slides as not
analysable, it being better to rely on historic performance of the slopes, possibly related to rainfall, to assess the
likelihood of sliding. There are numerous attempts to model the performance of such slopes using infinite slope
analysis, and modelling the net infitration of rainfall. These rely on assuming the conditions are as shown in figure
49(a).
They may in some cases attempt to model the partially saturated zone with some degree of theoretical rigor eg.
Fourie (1996), Pradel and Raad (1993).

RUN

R A IN FA LL
-O N

E VA P O TR A N S P IR ATIO N
GR O

S O IL

U ND
IN F L W A T E
R
OW

IN FILT R AT IO N

R UN
GRO

ROCK

RECHARG E
(= IN FILTR ATIO N IN
S IM P LE S T M O D E LS )

(a) Idealised slope

UND

W AT

ER P
IE Z O

-O F

MET
R IC
HEA
GRO
D
U ND
W
O UT
AT E
F LO
R
W

EVA PO TR A NS PIRATIO N

R AIN FALL

VE G ETATIO N

G RO U N DW AT ER
IN FLO W

IN TER C EP TIO N

R O O TS

F IS S
NON

N ETT R AIN FALL

( H IG C O L L U
H PE
V
R M E IU M
A B IL
IT Y

URE
DR
)
E S ID
- F IS
UAL
SUR
S
O
ED R
IL ( M
E DI
ES I
UM
DUA
P ER
L SO
MEA
I L (L
B IL I
OW
TY)
PE R
SE EPAG E
M EA
B IL I
TY )

SO IL P IPES

SA ND IN FILLED
JO IN TS

SE EPAG E

SE EPAG E

G RO U N DW ATER
O UT FLO W

C O M PLETE LY W E AT HE RE D
R O C K (LO W PE RM EA BILITY)
W E ATH ER ED , JO IN TE D
R O C K (M ED IUM PER M EA BILIT Y)

(b) Actual slope

Figure 49 : Hydrogeological conditions in shallow landslides on natural slopes


(a) Idealised slope (b) Actual slope.
The real situation is however far too complex to model successfully the real slope has roots, root holes, fissures,
desiccation cracks, variable soil properties with depth eg. higher permeability colluval soils underlain by lower
permeability residual soils and completely weathered rock; colluvium is commonly stratified. Alternating layers
noteably differing in permeability, create perched and pressurized bodies of ground-water which are frequent causes of
shallow slope failures; seepage inflow from the underlying rock etc; soil pipes (in some situations), and potentially
complex mechanics eg. collapse surface controlling failure rather than peak strength. Hence the calculations are at
worst simply misleading giving those doing the analyses a feeling of rigor in their analyses. At best, they can be
calibrated against real slope performance, but then it is questionable whether it is worth going through the steps of the
stability analysis, and whether it would be better simply to calibrate slope performance against rainfall, evaporation and
slope characteristics such as slope, catchment area above the slope, and geomorphological characteristics.
The use of monitoring of pore pressures in such slopes is very difficult (and essentially impractical except for research
purposes) because of the effects of partial saturation, water filled desiccation cracks, soil pipes, complex layering of
the soil over the rock, and the rapidly transient nature of the pore pressures.
(b) Natural slopes medium slides
By medium slides is meant slides of depth greater than 5 to 10 metres and up to several hundred thousand cubic
metres volume. For slides of this magnitude, it is reasonable to carry out limit equilibrium analysis of stability using
one of the several methods available. Numerical analyses are seldom warranted except in forensic studies (when
something has gone wrong and expenditure is not limited).
These slides are usually just as complex hydrogeologically as described for shallow slides, and prediction of the
pore pressures from the water balance is seldom attempted. It is necessary to instrument the slope with piezometers and
a rain gauge (or snow thickness and temperature) and monitor the slope for sufficient time to enable a relationship to be
developed between the piezometric levels and rainfall. Often the frequency of a piezometric level being exceeded is
determined from the frequency of the rainfall.

While this sounds simple enough, the reality is quite different because:
landslides are quite heterogeneous, and pore pressure response may vary significantly over the slide, or even in
piezometers quite close to each other.
the critical rainfall duration has to be assessed, as does the effect of antecedent rainfall. Often only 24 hour
rainfall data is available, when shorter duration rainfalls are more critical.
a long period of record is needed, Fell et al (1991) found that a semi-empirical model calibrated on 9 months
record with weekly readings (over several piezometers) gave very poor predictions. Even 3 years of records
was insufficient to get good correlation between modelled and recorded data. The rain gauge should be at the
landslide site, and should record short term eg. 15 minute rainfalls. Examples of such analyses are given in
Haneberg (1991), Okunushi and Okumura (1987), Nieuwenhuis (1991), Skempton et al (1989).
(c) Natural slopes large landslides
For slides of this magnitude (say 1 million m3 and above) limit equilibrium analyses will often be backed up by
numerical analyses to model the internal deformations. Consideration must be given to possible retrogression in the
landslide movement eg. movement beginning to reactivate at the toe, gradually extending to the crown of the landslide.
It must also be recognised that movement may occur over a wide range of factor of safety eg. movements at Downie
slide reduced from an average of 10mm/year prior to reservoir filling and remedial works, to 2 mm/year, after, even
though the factor of safety had been increased by 7% (Imrie et al (1991), Enegren and Imrie (1996)).
Pore pressures for analyses can only be determined by installing piezometers, taking account of the detailed
hydrogeology, including often complex, non-hydrostatic and compartmentalised pore pressure. These can be related to
rainfall and snowmelt, but as for medium slides, it is not a simple exercise.
(d) Cut slopes
The analysis of stability of cut slopes will usually be done by limit equilibrium methods, sometimes supported by
numerical analyses to determine stresses and deformations, and to assess the likelihood of progressive failure due to
strain weakening. Pore pressures in cut slopes are complex, with the problems associated with any natural slope,
compounded by the effects of the negative pore pressures due to unloading. These are discussed further below. For
existing cuts, the installation and monitoring of piezometers, and relating these to the hydrogeological conditions and
the rainfall evaporation and snowmelt which control the pore pressures, is the only way to get a reasonable estimate of
conditions. However it must be recognised that cracks and open joints etc may allow high transient pore pressures to
occur in the slope and these are likely to be missed by the instruments.
For new slopes it is usually impossible to accurately predict the pore pressures, and it is necessary to rely on an
understanding of the hydrogeology and precedent in similar slopes. The ideal situation is to monitor the pore pressures
as the slope is constructed, and install borehole drains or other measures to reduce the pressures if they are too high.
(e) Fills
The analysis will usually be done by limit equilibrium methods. Only occasionally, eg. in large dams, or in
investigation of unusual deformations which have been experienced by the slope will it be necessary to use numerical
methods. Where strain weakening and progressive failure is likely, numerical methods should be used.
For existing fills, which do not have drainage layers, the only reliable way to estimate pore pressures is to install
piezometers, and monitor them as detailed above for cuts and natural slopes. For new fills, where seepage may flow
into the fill from the slope on which it is constructed, the slope should be engineered with a drainage layer eg. as shown
in figure 50, to control the pore pressures. The drainage layer could be selected free draining rockfill, or a sand-gravel
drain or in some situations, a geo-synthetic.
4.1.3
The role of hydrogeologic numerical modelling
Hydrogeology frequently has a decisive role in slope stability assessment and has to be studied and evaluated in a
way commensurate to the magnitude of the problem. The modelling techniques are quite powerful and many programs
have also become very user-friendly (which unfortunately also increases the risk of incompetent usage). The treatment
of a landslide is a multi-disciplinary task. The participation of specialists in the field of hydro-meteorology and
hydrogeology may have to be considered, particularly for larger landslides. Subjects to be evaluated are the rainfall
regime, the infiltration rates (which in turn are related to interception, evapotranspiration, soil characteristics), the
aquifer parameters (storage, permeability, transmissivity) and the aquifer geometry. A further complication derives
from the variation of hydrogeological characteristics with the degree of saturation. This effect is particularly important
in soils where suction heads may contribute a notable proportion of stabilizing forces. The stress dependence of the
permeability in rock has been studied for a long time and can be included in some models. Similarly, the soil-water
characteristic curves are influenced by ambient stresses. The discussion above favors an observational approach to
hydrogeological phenomena, but only in the context of good quality hydrogeology and often in conjunction with
numerical modelling. The engineer or geologist in charge of the analysis should be aware of the limitations, mainly
related to the uncertainty of input.

R O A D FILL
B A S E O F C O L L U V IU M

D R A IN A G E O U T LE T

D R A IN A G E L AY E R

Figure 50 Drainage layer under road fill to control seepage pore pressures.
4.2
Modelling of Pore Pressures in Relatively Homogeneous Soil slopes
4.2.1 Unsaturated deposits
Infiltration of water in unsaturated soil deposits has been the object of several theoretical and numerical studies. Eg:
Lumb (1975(a)), Alonso et al. (1995), Collins and Znidarcic (1998) and Sun et al. (1998). Figure 51 presents numerical
results obtained for a slope in decomposed granite from Hong Kong. The saturated hydraulic conductivity of this
material is slightly higher than 10-4 m/s; the three sub-figures show the simulation of one-dimensional infiltration for
three constant rain infiltration rates. It can be seen that, even with the highest infiltration rate (80 mm/h = 2.2x10 -5 m/s),
the new infiltration zone remains with some suction and partially saturated, which is due to the fact that the infiltration
rate is smaller than the saturated hydraulic conductivity of the material. As the infiltration rate decreases, suction
increases.

a) Infiltration of 80 mm/h for 6 hours

b) Infiltration of 20 mm/h for 24 hours

Figure 51 : Results of transient infiltration


analyses for a granite soil from Hong Kong
(Sun et al 1998)

c) Infiltration of 5 mm/h for 96 hours

Sun et al. (1998) examined numerically and analytically one-dimensional infiltration in initially unsaturated ground.
They came to the conclusion that, for a water supply or rate of infiltration q1, the depth of the advance of the wetting
band with time t can be approximately defined as follows:
h = q1 t / n (Sr1 Sro)

(9)

in which n is the porosity, Sro is the initial degree of saturation and Sr1 is the degree of saturation in the wetting band. q1
is also equal to the hydraulic conductivity of the soil, k1, associated to Sr1, These two parameters, Sr1 and k1, are related,
through the soil-water characteristic curve and the permeability-matric suction function, to the matric suction, which
then takes the value u1. This later suction thus applies in the partially saturated wetting band. When the water supply is
equal to or larger than the saturated hydraulic conductivity of the saturated soil, ksat, infiltration is controlled by this
latter parameter and the excess water will become runoff. The maximum depth of the wetting band is still defined by
Eq. 9, but with q1 = ksat and Sr1 = 1.0, assuming that no air is trapped in the process. In these latter conditions, Eq. 9
becomes equivalent to the one proposed by Lumb (1975). Equation 9 also shows that the progression rate of the wetting
band depends on Sro, thus indirectly on antecedent infiltration and evapo-transpiration, an aspect which has been
observed in situ, in particular by Johnson and Sitar (1990).
As discussed in section 4.1.2, and observed by Lacerda (1989), Johnson & Sitar (1990), Montgomery et al. (1997)
and others, the development of pore pressures in a slope may not result from vertical infiltration in a homogeneous soil
only. Infiltration into the ground can be facilitated by the presence of cracks and holes made by plants or animals, flows
through more permeable soil layers and fractured bedrock. The hydrogeologic response of a hillslope to rainfall may
thus be extremely complex.
Because of the materials encountered and the complexity of the infiltration processes, the relation between climatic
conditions, in particular rainfall, increases of pore water with time and the possibility of landslide triggering cannot be
easily established, and can considerably vary from one place to the other. Brand et al. (1984) and Brand (1995) reported
studies which showed that the antecedent rainfall has no significant influence on the occurrence of landslides in Hong
Kong and that rainfall intensity is the triggering factor. Finlay et al (1997) however showed that, using data from
multiple rain gauges, antecedent rainfall did have some influence on the likelihood of landsliding in Hong Kong. De
Campos and Menezes (1992) showed that in Salvador, Brazil, there may be a delay of several days in between rainfalls
and landslide activity; in many places, it has been observed that landslide activity results from the combined effect of
antecedent rainfall and rainfall intensity. Figure 49 below shows the combined effect of the cumulative rainfall over the
two previous days and the hourly rainfall on the triggering of landslides in South-Korea.
Rainfall is however not the only cause for infiltration. In mountainous areas and in cold climate countries, snowmelt
provides a continuous supply of water over several weeks, generates an increase in pore pressure in the ground and is
often an important triggering factor (Jorstad, 1968; Wieczorek et al. 1989).

Maximum hourly rainfall (mm)

100
Nb events

80

Minor
Severe
Disaster

1-3
4 - 19
> 20

60
40
20
0
0

100

200
300
400
Cumulative rainfall

500

Figure 52 : Relationship between landslide events and rainfall in South Korea (Kim et al 1992)
4.2.2 Saturated deposits
Even in a saturated deposit, variations of the water table level or changes in pore pressure at the boundaries are not
always entirely reflected at depths in the deposit. This is associated to the fact that soils are compressible and it is
controlled by the coefficient of consolidation of the soil. As the strains involved are small, the coefficient of
consolidation has to be estimated on the basis of the small strain modulus (Leroueil, 2000). In soils with low coefficient

of consolidation, this phenomenon results in a magnitude of pore pressure variation at depth that are smaller than at the
boundaries. This has been observed, in particular, by Berntson and Sllfors (1984), Kenney and Lau (1984), Vaughan
(1994) and Demers et al. (1999). It may have practical implications when stability of slopes is considered. Quite often,
engineers base their stability analysis on the assumption of the highest possibly reached water table and associated
steady state conditions at depth. This latter assumption implies pore pressures in the deposit that may be higher than
they can possibly be, with the result that the factor of safety is underestimated when the slope is stable or that the
strength parameters back-calculated after a failure are overestimated.
Another practical implication of pore pressure variations in saturated soil masses is the stability of cuts in clay. As
indicated by Bishop and Bjerrum (1960), with the reduction in mean stresses due to excavation, there is a decrease in
pore pressure in the soil mass. This results in a short-term factor of safety at the end of excavation. With time, pore
pressures progressively go towards equilibrium and the factor of safety decreases towards the long-term factor of safety.
Figures 53 and 54 show the excavation at Saint-Hilaire, Qubec, and pore pressures observed in some of the
piezometers. The calculated pore pressures were obtained considering small strain shear modulus Gmax and measured
hydraulic conductivity, multiplied by a factor 2.3 to get the good fit shown on Figure 54. This indicates, however, that
pore pressure equilibration is controlled by a shear modulus close to the small strain shear modulus, a factor 2.3 being
relatively small when coefficients of swelling/consolidation are considered.
The time necessary for pore pressures to reach steady-state conditions varies with the consolidation characteristics
of the clay and with the geometry of the problem. Experience shows that this time can vary from days to million of
years (Leroueil, 2000). As indicated in Figure 55, for thick deposits of very low permeability clay shales (Pierre shale
in South Dakota (Neuzil, 1993) and Bisaccia in southern Italy (Fenelli and Picarelli, 1990) time for equilibrium could
be very long whereas at La Bosse-Galin, France (Blondeau and Queyroi, 1976), pore pressure equilibration was
following the excavation works. From field observations reported in the literature, Leroueil (2000) concluded that the
time for pore pressure equilibration is generally less than one year for excavations in soft clays.
Leroueil et al. (1990) and Leroueil (2000) indicate that in soft clays, the short-term factor of safety can be evaluated
on the basis of the vane shear strength; on the other hand, Chandler (1984) shows that in stiff (I L < 0.2) and often
fissured clays, the undrained shear strength of the clay measured in triaxial tests strongly overestimates the shear
strength mobilized during first-time failures in these materials, typically by a factor of two.
18

27

8m

32-11

24-13

24-13

23-8

Piezometer
Piezometer considered in Fig. 54

Figure 53 : Test excavation at Saint-Hilaire (Lafleur et al., 1988(a) and (b))


0
-10
Piezometer 32-11

-20
-30
0

Change in pore pressure u - uo (kPa)

-10

Measured
Calculated

-20
-30
-40

Piezometer 23-8

-50
-60
0
-10
-20
-30
-40
-50
-60

Piezometer 24-13
(see Fig. 50)

-70
-80

50

100

150
Time (d)

200

250

300

Figure 54 : Measured and calculated changes in pore pressure with time at Saint-Hilaire
(Lafleur et al., 1988(a) and (b) and Laflamme and Leroueil, 1999)

Time for pore pressure equilibration (years)

10 6
10
10

Pierre shale
Bisaccia (~ 100 m)

10
10

Isle of Sheppey (44 m)

London clay (6 - 12 m)
10
1
-1

10

-2

Saint-Hilaire (8 m)
Hede (5.5 m); Kimola (10 m)
Mexico City (4.5 - 8 m)
La Bosse-Galin (4 m); Bangkok (4 m)

10

Figure 55 : Time for pore pressure equilibration after excavation in clayey deposits (Leroueil 2000)
4.3
Limit Equilibrium Analysis of the Stability of Slopes
4.3.1
A General Review of the Methods
The theoretical basis of Limit Equilibrium (LEA) methods developed from classical rigid plasticity solutions,
utilizing Drucker and Pragers Upper Bound Theorem (eg. Chen, 1975).
If a stable, statically admissible state of stress exists for a given loading, the actual failure load is greater or equal
than the given load.
The theorem states the basic principle of LEA methods. The sliding body of an assumed shape is divided into slices
in plane strain, or columns in three dimensions. A state of stress is then determined by balancing an assumed
distribution of boundary stresses with the stresses induced by the weight of the elements.
Even though soils are neither rigid, nor perfectly plastic, thus invalidating the theorem, Limit Equilibrium methods
provide acceptable solutions in many situations. Their validity has been demonstrated by back-analysis of actual cases
and models, as well as by long experience in practical applications. Nevertheless, the Factor of Safety derived from
Limit Equilibrium methods should be considered more as an index of stability than as a physical parameter.
Several comprehensive reviews of slope stability analysis have appeared recently (e.g. Duncan, 1992, 1996(a), (b),
Morgenstern, 1992). Therefore, this review concentrates on only a few selected issues. From a large number of LE
methods in existence, the following are most frequently used in todays practice:
Bishop's Simplified Method (Modified Method in the USA) was developed for circular sliding surfaces (Bishop,
1955). It uses the vertical force equilibrium and the moment equilibrium conditions, but avoids the consideration of
horizontal force equilibrium. Its iterative solution is remarkably stable and efficient and produces highly accurate
results when compared with more rigorous methods. Fredlund and Krahn (1977) applied the method to non-circular
sliding surfaces, by including the moment of each normal force acting on the base of an element. This application
requires a suitable selection of the moment equilibrium center (Hungr, 1997).
Janbu's Simplified Method (Janbu, 1968), is very similar and equally efficient, but uses horizontal force equilibrium
condition instead of moment equilibrium. It was developed specifically for shallow non-circular rupture surfaces. It
tends to be excessively conservative for deeper surfaces and requires the use of a correction factor ranging from 1.0 to
1.1. The main advantage of the Janbu method with respect to Bishops is that the latter is inaccurate for problems
involving horizontal external loads. Both Bishops and Janbu Methods exist in three-dimensional forms, which behave
similarly as their 2D counterparts (Hungr et al., 1989).
The Morgenstern-Price (1965) and Spencer's (1967) methods are "rigorous" methods, which satisfy all equilibrium
conditions and allow for a certain degree of mobilization of internal strength. Their disadvantage compared to Bishops
or Janbu Methods is approximately 10 times longer computation time (not a serious constraint today) and, more
significantly, a tendency to diverge in certain cases. A 3D equivalent of Spencers Method was derived by Lam and
Fredlund (1993).
Sarma's Method (Sarma, 1973) is based on the assumption of the full mobilization of internal strength on assumed
(non-vertical) slice boundaries. The assumed level of internal strength and orientation of the internal shear surfaces
exerts a strong influence of the Factor of Safety. Consequently, the Sarma Method is most applicable to problems
where high internal strength can be safely assumed, such as landslides in structured rocks.
All the above methods, except for Janbu, produce similar results when applied to rotational (circular) sliding
surfaces (e.g. Duncan, 1996). The reason is that the various methods differ mainly in their treatment of the internal
strength of the sliding mass. Because rotational geometry requires little internal distortion, internal strength is not
relevant. The situation is different in case of non-rotational (compound) geometries, where significant internal

distortion must occur in order to make the sliding motion kinematically feasible. The following example shows a
comparison of the various methods for a non-circular surface (Hungr, 1997):
The example is a simple bi-planar sliding block (Figure 56), defined by a flat basal surface (Plane A) and a steeper
back-scarp (Plane B). Either of the two planes can be a weak surface. The second plane may be a stronger structural
feature, or it may be a rupture plane extending through intact material. A parametric study has been carried out,
varying the friction angle on Planes A and B between 10, 20 and 30 with no cohesion. Both dry conditions and porepressures determined by the piezometric surface shown in Figure 56 were examined. The resulting Factors of Safety
varied between 0.5 and 2.5 for the range of conditions.
Figure 57 shows a comparison between the various methods of analysis. The abscissa in this figure is the ratio, R,
between the friction coefficient on Plane A and that on Plane B:

R=

tan A
tan B

(10)

-m

The ordinate is a ratio between Factors of Safety determined by the various methods and that obtained by Spencers
Method, which is used as a reference. All the methods are shown to converge to the same value when R is low, i.e.
when the flat basal plane is weaker than the back scarp. On the other hand, the methods diverge by as much as 30% in
the case when the back scarp is weak and the sliding body is supported by the toe plane. Under these conditions there is
a significant difference even between Spencers and Morgenstern-Price methods.

Figure 56 : A bi-linear rupture surface used in the parametric study to compare


methods of stability analysis. The dashed line is the piezometric surface.

0.

Figure 57 : Comparison of the factors of safety obtained by four methods of analysis, for the bi-linear surface
shown in Figure 56. The factors are normalized with respect to the Spencers Method. The abscissa is the
ratio of strengths available on the two parts of the rupture surface (Hungr, 1997).

Janbus Simplified method produces a Factor of Safety which is consistently less than that of Bishops (with the
Fredlund and Krahn modification). The Janbu correction factor is required to compensate for this. All of these trends
persist in equal measure for dry conditions and with pore-pressure.
In practice, cases where R is low are fortunately more frequent than the opposite, as Plane A often follows a thin
weak layer in the stratigraphy, a bedding plane or a near-horizontal pre-sheared surface. Under such conditions, there is
not much difference between the four methods and the user is therefore justified in taking advantage of the high
efficiency of Bishops Simplified Method in two or three dimensions.
On the other hand, in cases where the strongest element of the sliding surface is at the toe, rigorous methods such as
Morgenstern-Price should be use. Examples of this are cases where the back scarp (Plane B) follows steeply inclined
bedding, a fault surface or similar and the sliding body is supported by a strong toe.
Donald and Chen (1997) describe a method based on the upper bound theorem of classical plasticity. The sliding
mass is divided into a small number of discrete blocks, with linear interfaces between blocks, and either linear or curved
bases to individual blocks. By equating the work done by external loads and body forces to the energy dissipated in
shearing, either a factor of safety, or a disturbance factor may be calculated. They claim the method gives an upper
bound estimate of factor of safety, so if combined with the Morgenstern Price method, which gives a virtual lower
bound solution, they give the range of the accurate answers that can be obtained.
Analysis of 2 and 3 Dimensional wedges, toppling and buckling are not considered here. They are covered in the
keynote lecture by E. Hoek, J. Read, Z Chen and A. Karzulovic (2000) Rock Slopes in Civil and Mining Engineering.
4.3.2
3 Dimensional Analysis
The Bishops Simplified, Janbu and Spencers Methods have been extended into three dimensions (Hungr et al.,
1989, Lam and Fredlund (1993). However, none of the available algorithms can account for the internal stresses
existing in a non-rotational and laterally asymmetric problem (Hungr, 1994). A true rigorous method that would
satisfy all the available equilibrium conditions has not yet been formulated. As suggested by Hungr (1994), a rigorous
three-dimensional method would require the definition of five spatially distributed inter-column force functions. The
solution for the Factor of Safety and five inter-column force inclination (l) coefficients would require six nested levels
of iteration. While modern computers could easily handle the numerical computation, the correct selection of the intercolumn force functions and possible convergence difficulties pose a daunting research problem.
Nevertheless, the use of three-dimensional analysis can be important, if for no other reason, then because it removes
the uncertainty connected with selecting the most representative cross-section. For example, the 2D Factors of Safety
for various sections of a lined landfill basin at Kettleman Hills, California (Figure 57) ranged from 0.81 to 1.35 (Seed et
al., 1990). The overall 3D Factor of Safety was 1.06 (Figure 58).

Figure 58 : A three-dimensional LE model of the Kettleman Hills landfill failure (Hungr, 1994).
The use of rules of thumb such as a 10% increase to compensate for the neglect of 3D effects, as suggested in
some textbooks, is not advisable, because although, as proven by Cavounidis (1987), the Factor of Safety of the critical
3D sliding surface always exceeds the critical 2D factor, the ratio between the two can vary within a range of 1.0 to as
high as 1.4 (Morgenstern 1992, Hungr, 1989). Danger lies in situations where the stability analysis is used to estimate
the strength of certain materials through back analysis. Neglecting a strong 3D effect in the back-analysis could result

in a serious overestimation of the back-calculated strength. However provided the geometry of the failure surface
remains the same for the back-analysis, and the forward analysis, this is not a major practical problem.
4.3.3
Limit Displacement Method
Recent advances in numerical solution of stress-strain problems have made it possible to apply what could be
termed a Limit Displacement method. Particularly suitable are Lagrangian techniques, which remain stable even with
very large strains (Itasca, 1992). The principle of the Limit Displacement Method is as follows: A sophisticated stressstrain model of the slope is set up, allowing for plastic yielding of the materials, controlled by material strength
parameters. An initial analysis is carried out, using the true strength parameters. For a stable slope, only small
displacements will occur. A second stage is run, with all strength parameters reduced by dividing with a factor, say 1.1.
Again, limited displacements will occur. The analysis is repeated with progressively increasing strength reduction
factor, until large displacements occur, signaling widespread yielding in the slope. The reduction factor used at this
stage is the Factor of Safety. The advantage of such an approach is that no search for a critical sliding surface is needed
and that effects such as progressive failure or even liquefaction can be simulated (e.g. Gu et al., 1993, Puebla et al.,
1997). The decisive disadvantage at present is the difficulty of estimating deformation properties and drainage
behaviour of many materials. For the time being therefore, the Limit Displacement Method is generally most suitable
for research purposes, or for studies of larger, more complex slopes while Limit Equilibrium methods remain prevalent
in practical use.
4.3.4
Some Common Problems
Some common problems in LE analysis which are often simply omissions, and or occur through lack of proper care,
include:
Use of circular rather than non circular analysis, where failure surfaces will clearly follow weak surfaces eg.
bedding surface shears.
Omitting to model cracks in the ground surface, and the water pressure which will often develop in the cracks
due to rainfall or surface water flowing into the crack.
Not modelling anisotropy of strengths eg. in fissured clays or stratified soil and rock.
These features can all be modelled with modern computer programs. If the program being used cannot model these
features, consideration should be given to purchasing a program which can.
A more subtle problem, which is an extension of the discussion on the different methods of calculation described
above, is where essentially translational sliding in rock is being modelled. Figure 59 shows such a slope, where the
rock is relatively strong, and with a few, widely spaced, near vertical joints AB, CD, EF (eg. sandstone or
conglomerate), with potential sliding on a weak layer (eg. claystone) OBDF. The driving force for instability is the
groundwater in the joints, and the gravity forces along OBDF.
E
C
A

ROCK
J O IN T S
B

0
W E A K L AY E R

Figure 59 : Limit equilibrium Analysis of a rock slope sliding on a weak layer


Use of a conventional computer program to analyse the stability of OAB, OCD, or OEF will over estimate the factor
of safety substantially even if the failure surfaces are pre-defined, because if one inserts a reasonable joint strength (say
c' = 0, ' = 40o), the program assumes this applies to the rock, and assumes slice slide forces will apply, equivalent to
say a wedge of rock with c' = 0, ' = 40o which will also drive the instability. This is difficult to overcome in the
computer programs, and it is better to analyse such cases using a wedge analysis (either by hand or a computer wedge
analysis).
4.4

Stress and deformation analysis methods


Duncan (1992, 1996) presents a history of the development of the methods of analysis with detailed tables listing
examples of the various methods of analyses.

Limit equilibrium analyses methods, (at least most of them) implicitly assume that: (a) the soil has a ductile (or
rigid-perfectly plastic) stress-strain behaviour; (b) the factor of safety is the same all along the slip surface; and (c) the
stress path up to failure is a very specific one. Tavenas et al. (1980) showed that the stress path implied in limit
equilibrium analyses are generally different from those followed in in situ conditions. The assumption of a ductile
stress-strain behaviour is appropriate at the reactivation stage where the soil is in residual conditions with strength
depending only on the effective stress. At the pre-failure and failure stages, most natural geomaterials display some
strain softening, so there is progressive failure which is not considered in limit equilibrium analyses. As a consequence,
such methods can be used only if average strength parameters, somewhere between the parameters at the peak and in
large strain or residual conditions, are considered. In these conditions, however, as indicated by Duncan (1992, 1996),
the value of F affords a useful index of the margin of stability for a slope.
Also, limit equilibrium analyses only consider the slip surface and not the stresses and strains of the soil around. To
get an insight of what happens in a slope before a first time failure, it is necessary to perform deformation analyses with
methods such as the finite element method. Pioneer work was performed in the mid-sixties but many studies have been
made since. Most of them examined excavations from an initially horizontal ground surface, where initial conditions in
terms of stresses (Ko) and strains (taken equal to zero) were clearly defined or constructed embankments, such as dams.
These include linear-elastic, multilinear elastic, hyperbolic elastic and elastoplastic and Elastovisco-plastic stress strain
relationships. Recent examples in the literature include Potts et al. (1990, 1997), Muir-Wood et al. (1995) and Kirkebo
et al. (1996) who used elastic-plastic models based on Mohr-Coulomb criterion or a slightly modified one. Potts et al.
(1990, 1997) introduced strain softening with two strength envelopes, one for peak conditions and one for residual
conditions. Cam-clay type models have also been used by Deschamps and Leonards (1992), Muir-Wood et al. (1995)
and Adachi et al. (1996). In this latter case, a strain-softening function was considered.
These numerical analyses confirm the process of progressive failure observed, in particular, by Cooper et al. (1998),
(see Figures. 33 and 34; Section 2.3.2). They also evidence a strong influence of Ko on this process. While in strongly
overconsolidated clays (say Ko > 1.0), zones of local failure always include the toe of the slope, they could be confined
below the crest of the slope in nearly normally consolidated clays (say Ko < 1.0), (Deschamps and Leonards, 1992).
Potts et al. (1997) examined numerically the effect of K o on the behaviour of a 10 m high and 3:1 cut excavated in
London clay over a three months period. Figure 60 shows the rupture surfaces predicted by the analyses and the
strength mobilized just prior to collapse. It can be seen that: the rupture surface becomes deeper as Ko increases from
1.0 to 1.75; only a short length of the rupture surface is at a strength between peak and residual; the rupture surface is
not completely formed at the time of collapse when Ko is equal to 1.0 or 1.25; with high values of Ko, the horizontal
rupture surface develops beyond the crest of the slope, as observed at Selborne (Figure. 35(d)).

Figure 60 : Rupture surfaces predicted by the analysis of 3:1 slopes, 10m high,
with surface suction 10kPa and varying Ko (Potts et al (1997)

Duncan (1996) in his State of the Art paper concluded regarding deformation analysis: each of these linear elastic,
multilinear elastic, hyperbolic, and elastoplastic relationships has its own advantages and limitations. Linear elastic
stress-strain relationships have the advantage of simplicity, and the limitation that they only model the behaviour of real
soils well at low stress levels and small strains. Multilinear elastic stress-strain relationships have the advantage that
they can be used to model any shape of stress-strain curve for ductile materials, and the limitation that they must be
developed on a case-by-case basis to approximate the particular stress-strain characteristics of the soils under
consideration. Hyperbolic stress-strain rrelationships have the advantages that they model nonlinear behaviour, and that
the parameters involved have physical significance and can be evaluated using the results of conventional triaxial tests.
They have the limitation that they are inherently elastic, and do not model plastic deformations in a fully logical way.
Elastoplastic, and elastoviscoplastic stress-strain relationships have the advantage that they model more realistically the
behaviour of soils close to failure, at failure, and after failure. They have the limitation that they are more complex.
Comparisons of the results of finite-element analyses with field measurements have shown that there is a tendency
for calculated deformations to be larger than measured deformations. The reasons for this difference include: (1) soils
in the field tend to be stiffer than soils at the same density and water content in the laboratory because of aging effects;
(2) average field densities are higher than the specified minimum dry density, which is often used for preparing lab
triaxial test speciments; (3) samples of inplace materials suffer disturbance during sampling, and are less stiff as a
result; (4) many field conditons approximate plane strain, whereas triaxial tests are almost always used to evaluate
stress-strain behaviour and strength; and (5) two-dimensional finite-element analyses overestimate deformations of
dams constructed in V-shaped valleys with steep valley walls.
Duncan (1996) concluded (as had Vaughan (1994), who discusses several cases), there is much to be gained from
the analyses, in the understanding of mechanisms, if not the absolute stresses and deformations.
The authors experience in this area is limited to that of user rather than doer of the analyses. From this it is
clear that it is difficult to model the deformation of slopes with any degree of precision, in particular because we do not
have any origins for strains for natural slopes. The ability to numerically model generally is greater than the ability to
provide accurate material properties.
There are however often more basic reasons for poor modelling, including not constructing (or numerically
building) the slope in steps correctly, assuming drained conditions will apply when in fact undrained behaviour is
actually present and not paying enough attentions to the stiffness of the layers (or in particular relative stiffnesses), so
the stress distributions and deformations are poorly modelled.
Few model pore pressures accurately, or allow for cracking, softened zones around cracks, crack water pressures, or
in many cases, strain weakening. Given that these factors often control the stresses and deformations in slopes, it is not
surprising the modelling is not good.
4.5

Probabilistic analysis of the stability of slopes


The use of probabilistic or reliability methods for assessing the stability of slopes is not common for other than mine
pit slopes. This is despite a large body of research and publication into the topic, which have brought the state of the art
to a relatively high degree of sophistication.
A critical review of the state of the art as given in Mostyn and Fell, (1997) which draws on Mostyn and Li (1993)
and Mostyn and Small (1987). Other general references include Christian et al (1992), Wu et al (1996), Whitman
(1984).
There are merits in carrying out probabilistic modelling of the stability of more important slopes provided the
modelling takes account of the important variables, including geometry, pore pressures, shear strength and the effects of
spatial correlation, autocorrelation of properties, correlation between failure surfaces, and potential model bias. It is
clear that it is difficult to provide the data for such analyses, so a considerable degree of judgement is going to be
applied. As a result the answers are not going to be accurate in absolute terms, but will be more reliable in relative
terms eg. to compare different design approaches, different remedial works. The process is also likely to highlight the
areas of uncertainty. It should be noted that the analyses generally stop at the estimation of the probability of a factor of
safety of 1.0 being achieved (or expressing the same in reliability terms). For many situations this will not be sufficient,
and the further conditional probabilities relating to travel distance, and temporal probabilities of persons for example,
need to be considered. As for all analyses, it is essential the mechanics of failure are properly modelled. For example,
there is no point carrying out an analysis issuing effective stress parameters, if undrained strengths control the stability,
or using peak effective strengths in an analysis, when a collapse surface will control the initiation of failure.
4.6
Use of Monitoring to Predict Performance of Slopes
4.6.1
Monitoring instrumentation
Monitoring of landslides serves multiple purposes. It constitutes an essential component of the investigation of a
landslide (investigative monitoring), it helps to control the implementation of stabilizing works and it is used to observe
the performance of the slope in the context of risk management and, eventually, to warn of impending emergencies.
Therefore, monitoring should start in the early stages of an investigation program and it may have to continue
indeterminately.

Monitoring typically covers


magnitude, rate, location and direction of deformations
pore pressures and piezometric levels, hydrogeological and hydro-meteorological parameters
seismic accelerations (only occasionally).
Stresses are rarely monitored in slopes but forces on slope support eg. anchoring, may have to be kept under
observation.
In selecting the method of monitoring and the type of instrumentation, the required accuracy and durability in
relation to the available time and the expected rate of deformation has to be considered. For instance, a borehole
inclinometer will precisely detect minor deformations but its lifetime will be quite limited when it traverses an active
failure surface. Alternatively a shear plane indicator (downhole slope extensometer) will not give quite the same
information and not with the same sensitivity and accuracy but it may endure large displacements.
Dunnicliff (1995) describes instrumentation and methods for slope monitoring. An ICOLD committee on Reservoir
Landslides has recently updated and complemented this list which is presented in Table 14.
In most cases, monitoring will consist only of surface deformation using survey and inspection for cracking and
other deformations, borehole inclinometers, piezometers, drainage flow (if applicable) and rainfall, (or snowfall and
snowmelt).

TABLE 14 : INSTRUMENTATION AND METHODS FOR SLOPE MONITORIN


(ADAPTED FROM DUNNICLIFF 1995)
Parameter
Surface deformation

Subsurface deformation

Groundwater pressure

Stresses in slope reinforcement


Drainage flow (from borehole drains, trench drains,
adits), springs and surface runoff
Climatic conditions

Instruments/Methods
Surveying methods, including GPS
Crack gauges/surface extensometers
Tiltmeters
Multi-point liquid level gauges
Photogrammetry
Satellite images
Remote video
Inclinometers
Simple borehole deformation measurements
Fixed borehole extensometers
Slope extensometers
Shear pin indicators
In-place inclinometers
Multiple deflectometers
Acoustic emission monitoring
Time-domain reflectometry (coaxial cables)
Pendulum
Standpipe piezometers
Vibrating wire piezometers
Pneumatic piezometers
Multi-point piezometers
Load cells
Strain gauges
Bucket and stopwatch
Water level behind weir(s)
Pipe flow meters
Rainfall gauges
Thermometers
Barometric pressure gauges
Evaporimeter
Tensiometers
Snowpack depth and density gauges

Rainfall duration and intensity may be used alone as an indicator of the likelihood of incidence of landsliding. Such
systems are used in Hong Kong (Brand 1995) and in Rio de Janeiro (dOrsi et al 1997) and in local areas eg. parts of a
railway susceptible to sliding. They are usually set to trigger on a conservative basis, so the population at risk can
avoid hazardous areas.

4.6.2
Observational method and its limitations
In the investigation and treatment of landslides, two problems are commonly experienced:
(a) Even if time and funds are available for investigation and analysis, a considerable margin of uncertainty
remains in respect of:

the strength parameters and their development with time and strain,

the accuracy in the conceptual kinematics and the geometry of the geotechnical model of the landslide,

hidden geological details which may significantly influence the performance of the slope,

the possible range of hydrogeological and seismic loads,

the velocity and travel of post failure movements.


(b) Given the uncertainty, the need for measures to be taken against exceptional risks and the most suitable way of
dealing with them remain poorly defined.
This situation may lead those responsible to either opt for a high factor of safety or to disregard the exceptional risk,
but, as pointed out in Peck (1969), the first method is wasteful; the second is dangerous. In some cases, eg. large
natural landslides, or mine pit slopes, high factors of safety are impractical. In these situations, the observational
method offers a solution for the dilemma. Introduced by Terzaghi and Peck, it is an accepted approach in geotechnical
engineering, as formalized in Eurocode 7 (1997). In close agreement with the 1969 Rankine lecture by Peck (Peck
1969), Eurocode 7 defines four requirements for the observational method:
1. The limits for acceptable behaviour must be established.
2. The range of possible behaviour shall be assessed and an acceptable probability shall be demonstrated for the
behaviour to stay within acceptable limits.
3. A plan of monitoring shall be implemented to verify compliance with the defined limits. Response time of the
instruments and reading intervals should be adequate to allow contingency action to be taken.
4. A plan of contingency actions, to be adopted in the event of adverse developments, must exist.
Several conditions must be met in order to cope with these requirements. The monitoring system must be capable to
detect the significant warning signs accurately and reliably. This involves verification, maintenance and possibly also
successive modifications of the monitoring system. An efficient organization and clearly defined routines for recording,
reducing and evaluating the data have to be established. Although criteria for contingency action have to be specified in
advance, the nature of the problem will always call for competent judgement. The review of the monitored
performance of the slope has to assess the adequacy of instrumentation and procedures for monitoring, the concepts of
the contingency action and the criteria for their implementation.
As a large number of case histories demonstrates, the observational method offers substantial benefits in the
treatment of high risk slope stability problems, but the observational method has also pitfalls and limitations. The
possibility of sudden, brittle failure constitutes one of the most serious limitations of the methods, so it is essential that
the mechanics and mechanisms of sliding are clearly understood. The observational method is not suited for conditions
which limit the possibilities for contingency action or where contingency action cannot be implemented sufficiently
fast. The method is prone to fail if the monitoring installation does not cover the relevant parameters and processes or if
the instruments give wrong or inaccurate information.
The main limitation relates to the frequency of readings required to adequately monitor a slope. There is little value
in relying on annual inspections and readings of deformations and pore pressures as a risk management tool, when most
slopes have a response time to the driving forces (rainfall, snowmelt) which is in terms of hours, days or months
(depending on the slope). At best, such annual readings may identify slowly deteriorating conditions eg. where seepage
through a dam is slowly wetting and softening the earthfill.
4.6.3
The monitoring of slope movements prior to failure general issues
There are some general principles involved in using the measurement of deformations of slopes to predict
performance:
(a) slopes which are approaching failure conditions are assumed to display creep characteristics, shown in figure
30).
The strain rate occurring in the primary creep period has been found to decrease with time, following either an
universe power law, or exponential law. Strain during secondary creep is linearly proportional to time, and strain
during tertiary creep follows either a power law, logarithmic law or exponential law (see 4.6.4).
(b) Pre-failure movements include all the movements that occur before a first-time failure. They can result from a
combination of phenomena: elasto-plastic deformations associated with changes in effective stresses; viscous
deformations; softening, and strains and displacements associated with progressive failure.
A characteristic of pre-failure creep movements is that their rate varies with the stress conditions in the slope, and
often with the seasonal rainfall or snow. According to viscous models, as the stress path gets closer to the limit state
curve (or to the peak strength envelope) of the soil, creep deformations develop at increasing strain rates, as indicated in
Fig. 29(b). As, with the seasonable variations of pore pressures, the effective stress conditions fluctuate between limits
such as HW, when pore pressures are high, and LW, when pore pressures are low, creep rates vary accordingly from
high values (when at HW) to small values (when at LW) Such variations have been observed in slopes in eastern

Canada clays by Mitchell & Eden (1972); the creep rates observed during the spring season, when the water table is at
its higher level, were 8 to 33 times larger than the average creep rates measured over a 3 year period.
(c) It is essential to assess whether the slope is experiencing first time sliding or whether it is an active, or
potentially reactivated slide.
For first time slides there is a potential for loss of shear strength due to strain weakening. The deformations required
to reach peak strength may be quite small eg. only millimeters in single rock joints (Barton 1982). Although for a
rock mass larger deformations usually occur before failure. For soils, this is sometimes measured by the brittleness
index IB which is discussed in Section 2.1.1.
(d) All constructed slopes deform when they are built, and continue to move with time. Those which are not
highly stressed, behave in a primary creep behaviour (see figure 30) and the movements will become so slow as to be
imperceptible. The question arises as to how much movement can you expect and what is abnormal?
Hunter and Fell (2000) have gathered good quality data for central core earth and rockfill dams. The settlement
during construction of the embankment core is generally 1% to 3% of the height of the dams, for dams up to 100m high.
Post construction settlements are generally 0.2% to 0.6%. Dams with poorly compacted rockfill shoulders generally
experience larger (0.6% to 1%) post construction settlements. In some cases where these are linked to large lateral
movements, localised yielding in the core appears to have occurred.
Figure 61 shows plots of settlement of several central core earth and rockfill dams.
0.0
Dam G
Dam F
0.2
Dam J

Settlement/Dam Height (%)

Dam H
0.4

Dam D
Dam A - Type 5, clay co re

Dam B

0.6
Dam B - Type 5, clay co re
Dam A

Dam C - Type 5, clay co re


Dam D - Type 5, clay co re

0.8

Dam F - Type 5, clay co re


Dam G - Type 5, clay co re
Dam C

1.0

Dam H - Type 5, clay co re


Dam J - Type 5, sand/silt core

1.2
10

100

1000

10000

100000

Tim e (Days)

Figure 61 Normalized settlement of the embankment crest, settlement/dam height,


with log time for some central core earth and rockfill dams (Hunter and Fell 2000)
It can be seen that the plots of settlement are basically linear indicating that the long term settlement is in the form of
a primary creep.
There are however some interesting features in the plot:
Dam C has a change in slope at around 2,000 days, and a major change at 12,000 days. These changes coincide
with large drawdowns in the reservoir level, and were accompanied by cracking of the dam crest, and lateral
upstream displacements. Detailed investigations and monitoring have demonstrated that the dam had a factor of
safety of 1.0 for small upstream slides, and a shear surface consistent with this was identified with borehole
inclinometers. The inclinometers and surface displacements showed that a deeper wedge of the core had also
dropped towards the upstream despite the factor of safety for deeper surfaces being 1.3. This was a result of
the low modulus dumped rockfill providing little lateral support to the core. The dam core was extensively
cracked, and softened by surface infiltration along the cracks. The change in behaviour after 2,000 days might
be due to the onset of cracking and softening.
Dam A shows a relatively large continuing rate of settlement. The displacement versus log time plot of the crest
settlement measurement point shows a similar large continuing rate of movement. The reason for this, and its
relatively abnormal total settlements during and after construction, are being further investigated, but appear to
reflect weakening of the core due to the large displacements during construction as was the case for Dam C.

Dam B shows a marked change in performance after a major reservoir drawdown at around 3,500 days. This
change in behaviour coincided with a localised increase in strain in the Internal Settlement Gauge approximately
10m below crest level, so again may reflect deformation and weakening of the core.
The horizontal movement versus log time has also been plotted, but has not been included here. It shows that most
of the non-recoverable movement occurs early in the life of the dam. For Dams A and C particularly, significant nonrecoverable movements also occurred on drawdown events much later in life, after 10,000 days.
The change in behaviour for these cases can be directly related to a physical change cracking and softening of the
dam core by water infiltration.
(e) In slopes which experience progressive failure, deformations, and the formation of shear surfaces, may begin
locally and progress to eventually form a complete slide surface. Cooper et al (1998) and Bromhead et al (1998)
present details of the progressive failure of the Selborne trial cut in overconsolidated clay (see section 2.3.2).
4.6.4
Prediction of time to failure and critical velocities for first time slides
If one is monitoring the deformation of a slope which develops tertiary creep characteristics, it is useful to have
some guidance on the likely future behaviour of the slope, in particular, how long it will take to reach the collapse
conditions and at what velocity is collapse imminent. Saito and Uezawa (1961) and Saito (1965) developed the first of
these methods.
Saito uses a phenomenological approach to estimate the time to failure (tf) from strain measurements. Using
laboratory tests, as well as with data recorded on landslides in soils, he found the following relation between total creep
rupture life (combining secondary and tertiary creep) and secondary steady strain rate (or alternatively transient
minimum strain rate, Saito (1980b):

= b / (t f t )
where

(11)

strain rate [eg. 10-4/min], b = constant

or, more specifically:

log t f [min ] = 2.33 0.916 log ( ) 0.59

(12)

ie. the confidence band of 95% for the estimate of log (tf) has a width of 1.18 which means approximately a factor of 15
for tf. (Brittle materials will have shorter rupture life than the above equation indicates).
For the tertiary creep range Saito (1980a) proposes an empirical equation:

((

)(

0 = A log t f t 0 / t f t
where

))

(13)

reference time and = strain or

((

)(

x = xA log t f t o / t f t

))

(14)

The latter equation can be solved graphically for tf by plotting log (t f t ) against the displacement x. If tf is
selected correctly, the graph will plot as straight line. As a first approximation, tf can be estimated from minimum strain
rate. Another method (Saito 1980a) estimates rupture life by reading the time for two successive intervals of equal
displacement x from the time-displacement graph and solving the above equation for tf. Saitos methods also resulted
in satisfactory predictions for landslides in weak rocks (Saito 1980a). Kawamura (1985) describes creep by the
following modified equation:
dt 1
= = A(B t )t
dx

(15)

where A, B are constants and t = time


This is more universally applicable because it covers both the cases of accelerating (divergent creep) which leads
to failure, and progressively attenuated movement (convergent creep). If A and B are >0, the creep becomes
divergent, for A<0 and small values of B, the movements will abate. Kawamura proposes to solve for the constants A
and B by reading in the time-displacement graph the time abscissa corresponding to a succession of equal intervals of
xi and fitting a linear equation to

(t i +1 t i ) / t i
where
and

= f (t i +1 ) = t i +1

(16)

= e (AB )
= /B

Continuing experimental studies by Fukuzono (1985) lead to the following relation between acceleration and creep
(as formulated by Voight, 1988)
d2x
dx
= A
dt 2
dt

(17)

where x = displacement, t = time A, are empiric constants


is typically: 20.3
which integrates to

1
dt 1
= = [A( 1)] 1 t f t
dx

1
1

(18)

=
time from beginning of record to failure, and
tf
v
=
rate of displacement
or in a condensed version (Azimi & al., 1988)

where

= / (t f t )

(19)

where , B are constants and B = 0,58


which differs notably from the earlier empirical equation given by Saito.
Voight (1988) points out, the method applies to first-time and reactivated slides in rock and soil and to toppling
failures, provided that stress remains constant. However the authors believe the methods should generally be applicable
only to first time slides, since the mechanics of reactivated slides are potentially quite different. The time to failure, tf in
the above equation, is estimated by graphical methods, for instance by plotting the logarithm of inverse of the
displacement velocity against time:

log() = C1 C2 log t f t

(20)

where C1, C are constants


In a procedure similar to that of Kawamura (1985) Azimi et al. (1988) propose to subdivide the smoothed timedisplacement graph into successive constant displacement increments xi and reading the time ti corresponding to the
end of each increment. Plotting the time corresponding to the ith increment, ti, against the time for the preceding
increment, ti-1, renders a straight line which intercepts the time ti=ti-1 at the time tf.
According to Azimi et al (1988) the procedure offers a convenient method to estimate the rupture life from tertiary
(accelerated) creep records of landslides. The equation given by Kawamura has the advantage that it distinguishes
between convergent, progressively attenuated creep and tertiary, accelerating movements.
Varnes (1982) presented a valuable overview paper on creep of geomaterials, and compared results with creep
testing of other materials such as metals and ice. He attempted to model the tertiary creep stage of a number of failures
(including laboratory and field cases), using exponential, power or Saito laws. Varnes (1982) concluded:
1. The most appropriate function for analysing the primary creep stage is the power law. Primary creep analysis of the
laboratory cases indicated they followed either logarithmic or power laws;
2. Most of the field cases exhibited tertiary creep of pure Saito or generalised Saito form (or a combination of both).
Eight of the cases examined follow a pure or generalised Saito form for at least part of the time, one case is of
exponential form and two follow a power law expression; and
3. In a number of cases the mathematical form of the tertiary creep curve changed. Often there was a change in form
immediately prior to collapse (the overall process, such as pure Saito may remain the same, but the constants
involved in the form of the curve may change).

Glastonbury (1999) analysed case studies of mine pit slope collapse, and combined these with case studies (mostly
mine pit slopes) from Zavodni and Broadbent (1980), Ryan and Call (1992) and Salt (1988) to produce figure 62.
Glastonbury (1999) concluded that:
1. For a given displacement rate at onset of tertiary creep, the expected period of tertiary creep can vary by
approximately 1.5-2.0 orders of magnitude; and
2. Cases with similar failure mechanisms and geology may be expected to lie in the same region of the graph, as
indicated by the Zavodni & Broadbent cases. These cases represent sliding failures in porphyry copper mines.
The authors see merit in plotting the data to assess whether the slope is experiencing primary, secondary or tertiary
creep. If tertiary creep behaviour is observed, the slope should be assessed to be in a state of marginal stability, and
potential collapse, until proven otherwise. The authors do not recommend too great a reliance on predictions of times
for failure as they are clearly uncertain.

Figure 62: Relationship between displacement rate at onset of tertiary creep,


and duration of tertiary creep(Glastonbury 1999)
Several authors have given opinions on what constitutes a critical velocity the slide velocity when failure is likely
to occur within a short period (days or hours?). These are summarized in Table 15

TABLE 15 : OPINIONS ON CRITICAL VELOCITY INDICATIVE OF IMMINENT COLLAPSE


Author

Suggested critical velocity

Situations

Salt (1988)

50mm to 100mm/day

Large natural slides

Wylie and Munn (1978)

720mm/day

Toppling failure in coal mine pit

Ryan and Call (1992)

12mm/day marked initial tertiary


creep

Mine pit slopes

50mm/day occurred at least 48 hours


prior to collapse. The velocity at 24
hours prior to collapse was 1.7 to 2.3
times that at 48 hours
Martin (1993)

10mm/day to 100mm/day suggests


progressive failure with acceleration
to collapse

Mine pit slopes

Zavodini and Broadbent (1980)

50mm/day, collapse likely within 48


hours

Mine pits

The authors caution against the use of these critical velocities, particularly for situations different to what they were
derived. Logically the critical velocity should be dependent on the scale of the slope, the mechanics of sliding and the
geology. In any case, from a risk management viewpoint, the consequences of failure should also be considered.
The mechanisms which lead to tertiary creep behaviour are not considered in much of the above discussion. Most of
the approach is empirical, and simply finding a mathematical equation which fits the data points. In many cases the
cause of the acceleration in slope movements will not simply be due to the shearing of the soil or rock under constant
stress conditions, but to changes in pore pressures, softening of the slope by seepage or other physical process.
Hutchinson (1987) lists some other mechanisms which can cause large, rapid displacements. He lists these in terms
of landslides on pre-existing shears, but they would also apply to first time slides. They are:

entry of free water into cracks eg. from overland flow from rainfall and from streams flowing over the slide
surface

pipes, dams, sewers crossing the slide area, and discharging their contents into the slide mass when the slide
moves

choking of groundwater seepage outlet eg. by freezing, progressive clogging due to internal erosion of the
slope or blockage by displacement of superficial lobes

drawdown and impoundment of reservoirs

head loading, and toe unloading of slides

failure of retaining structures.


These do not include strain weakening which is discussed elsewhere.
4.6.5
The velocity of active landslides
Due to the viscous nature of soils (Bracegirdle et al., 1992; Leroueil and Marques, 1996), the rate of movement, ,
does not abruptly vary from 0 when the factor of safety is larger than 1.0 to infinity when it becomes equal to 1.0, but
progressively increases with the applied shear stress level. Because the particle arrangement does not significantly
change after large displacements, the rate of displacement is not influenced by the amount of displacement, time or
history of the movement. This is not necessarily the case in rock slides, Vulliet (1986), who reviewed the literature on
the topic, suggested that should be expressed as follows (see also Vulliet and Hutter, 1988):

= F ( n , )

(21)

in which F is a function of the normal effective stress, n , and the applied shear stress, . It can take several forms,
one of then being the power law:
F = A n 1 / ( n tan r )n

(22)

There is evidence (Cartier, and Pouget 1988, Pouget and Livet 1994, Morgenstern 1995, Hutchinson 1988, Skempton et
al 1989, Niewenhuis 1991, Leroueil 2000), that the rate of displacement of active and regularly reactivated landslides, is
related to the factor of safety. Figure 63 shows the relation between the rate of displacement of an inclinometer through
the slide surface at a depth of 6 metres for a test embankment on an already unstable soil slope.

Figure 63 : Relation between rate of displacement (v) at the toe of the Salldes
embankment and factor of safety (F) (Leroueil 2000 after Cartier and Pouget 1988)
It can be seen that a 0.1 change in factor of safety approximates to an order of magnitude of change in rate of
displacement and that significant movements ( 1mm/day) were occurring at a factor of safety of about 1.1. Leroueil
(2000) points out that it is a small change in rate of movement, and suggests, on the basis of field observations (Bertini

et al 1986, Hutchinson 1988, Azimi et al 1992), and laboratory results, that typically, the rate of displacement decreases
by two orders of magnitude for a factor of safety change of 0.05.
For very large active landslides, the rates of movement also vary with factor of safety, but are relatively small. For
example, for the Downie landslide (Enegren and Imrie (1996), a change in factor of safety of about 0.07 to 0.11,
resulted in rupture surface movements (as measured in inclinometers) decreasing from 3 to 10mm/year to 0.75
2mm/year (about 0.17 to 0.25 of the original rates). For the No.5 Creek slide, Cromwell Gorge (McFarlane and Jenks
1996), a 0.09 change in factor of safety reduced average rates of movement from 13mm to about 3mm/year.
It is clear that many factors would affect these figures, including the scale of the landslide, the geology and slide
mechanism. It is also apparent that surface measurements may be misleadingly high compared to measurements at the
surface of rupture, possibly due to differential or surficial sliding superimposing on the main slide movement. It does
appear however that provided the means to measure the displacements are sufficiently sensitive, movements can be
detected at the early stages of slide reactivation. What velocity one should accept as representing a factor of safety of
1.0 is not so clear, particularly for landslides, such as the Downie, No. 5 Creek and Sallades where the rate of
movement is not very sensitive to factor safety. Often the only reliable approach in these cases will be to aim to achieve
a required increase in factor of safety with remedial action if that is required.
It would be unwise on the current understanding to extrapolate the rates of movement versus factor of safety of
active landslides into the below a factor of safety of 1.0 region ie. for example, assume that a factor of safety of 0.9
might only result in a rate of movement say 5 times that it is experiencing (at an implied factor of safety of 1.0).

5.

ANALYSIS OF POST FAILURE VELOCITY AND TRAVEL DISTANCE

5.1
Post Failure Behaviour of Sliding Movements
5.1.1
Sliding Block Models
The simplest type of analytical approach for post-failure motion is the sliding block model, which describes the
landslide as a dimensionless body moving down the profile of the path (figure 64). Its movement is controlled by a
single force resultant, representing the gravity driving force as well as all movement resistance. The correct application
of this "lumped mass" approach is to consider the movement of the centroid of the displaced mass (Banks and Strohm,
1974). However, following a trend first established by workers in snow avalanche dynamics (Voellmy, 1955, Perla et
al., 1980), many of the lumped mass models have been applied in terms of a hypothetical block travelling from the crest
of the source to the extreme distal point of the deposit.
Banks and Strohm (1974) carried out analyses of wedge rockslides in order to estimate the potential for landslidetriggered waves in a reservoir. They assumed the resisting force in the block to be a purely frictional term, dependent
on the normal component of the slide weight. This approach is probably correct for small scale rockslides of limited
displacement, which do not disintegrate during motion. It always predicts high velocities. As shown by Koerner
(1976), the line connecting the centers of gravity of the block before and after the slide represents the kinetic energy
head. The elevation of this line above the slide path equals v2/2g, where v is the velocity of the center of mass. The
ratio of the vertical to horizontal displacement of the center of gravity of the block equals the friction angle used in the
analysis.

X
Y

Figure 64 Sliding block model. The distance X is the kinetic energy head (v2/2g)
for a frictional model, Y for a model with viscous rate dependence.
Slides with long displacement, even if they do not break-up, become distorted and change shape in plan and profile.
British Columbia Hydro (1981, unpublished) analysed potential landslides along the shoreline of a reservoir using a
coupled series of blocks deforming in a user-prescribed fashion and controlled dynamically by a single constant
frictional resistance parameter. Romero and Molina (1974) derived dynamic equations for a series of connected vertical
slices, as frictional blocks. A similar analysis was carried out by Hungr and Evans (2000). As shown in figure 65, this
is a two-dimensional dynamic analysis of motion at a typical cross-section of the Vaiont Slide (Hendron and Patton,
1985). The analysis was carried out using the numerical solution of Hungr (1995), modified so as to use Cartesian
coordinates and vertical element boundaries on account of the great thickness of the slide. The frictional model was

used, assuming that the average friction angle on the sliding surface was 10 with no pore-pressure, which therefore
gives a lower bound for the friction angle. The analysis simulates well the displacement and velocity of the actual slide,
as reported by Mller (1964) and other workers. The question is why has the friction angle during motion reached such
a low value. Hendron and Patton (1985), carried out a similar analysis, and explained the low dynamic friction angle by
firstly, the presence of pre-sheared clay filled seams and secondly by a strength reduction due to frictional heating. Tika
and Hutchinson (1999) subsequently measured a certain velocity-dependent strength reduction on samples of Vaiont
clay tested in a ring shear apparatus. The analysis also ignores any brittle behaviour of the lateral restraints to the slide
(Hendron and Patton 1985).
a
E L E V A TIO N (m )

1 200
1 100
1 000
9 00
8 00
70 0
60 0
0

2 00

4 00

6 00

8 00

1 000

1 200

1 400

1 600

1 800

2 000

D IS T A N C E (m )

V E LO C ITY (m /s)

b
40
30
20
10
0
10 00

1 200

1 400

1 600

D IS T A N C E (m )

Figure 65 : A two-dimensional dynamic analysis of a section of the Vaiont slide, using an average
friction angle of 9o on the rupture surface (bulk friction angle, including pore-pressure effects).
(a) Profiles of the moving mass, plotted at 5 second intervals; b) velocity of the front and tail of the
slide mass (Hungr and Evans, 2000).
While large, deep-seated landslides may exhibit velocity dependent reduction of dynamic shear strength, the
opposite appears to be true for most slides and rotational slumps in dilatant, over-consolidated clays and weak rocks,
such as shales. These landslides rarely exceed rapid velocity (3m/min) and usually remain in the moderate range
(metres per hour or metres per day), even during the first time failure. Yet, any analysis using a frictional block model,
would yield velocities measured in metres per second, i.e. in the extremely rapid range. Consequently, some viscous
rate-dependent strength increase effect must exist in such dilatant plastic materials. Nisbet (1973) attempted to backanalyse slumps in over-consolidated clays using rate-dependent residual strength. However, he was unable to find
sufficient data concerning the time dependence of strength at higher shear strain rates. Wedage et al (1998) used a
laboratory-derived rate-dependent residual strengths to predict slow creep deformations of a tailings embankment, with
encouraging results. There is a difficulty in applying this approach to landslides with large displacement, as there is
very little known about the structure and rheology of shear zones. A shear zone in overconsolidated clay or weak rock
may be tens of centimeters thick and comprising several primary and secondary discrete shear planes, as well as plastic
deforming material. The rheological behaviour of such a shear zone cannot be studied in small-scale laboratory tests
and no accepted technique exists for extrapolating the results of such tests to the field scale. Consequently, we are
presently not able to make predictions concerning the post-failure motion of this type of landslides using this type of
approach.
5.1.2. Conservation of Energy Model
Khalili et al (1996) present a simplified method for estimating failure induced deformations in embankments which
remain essentially intact. The method is based on the principle of conservation of energy and the equation of
equilibrium, and like the sliding block model, assumes the energy loss (potential energy) is due to frictional forces
acting on the slip surface. It is based on figure 66.
The method considers two cases rapid and slow failure, which give upper and lower bound estimates of
displacement as
i - f =
which can be approximated as

cos f cos i
FS residual sin i

(23)

f = 2i (FSresidual 0.5)

(24)

for the rapid model, which gives an upper bound estimate and
Sinf = FS residual Sini

(25)

for the slow model which gives a lower bound estimate

Figure 66 : Failure model (Khalili et al 1996)


The method showed quite good correlation with the performance of embankments on soft clay (where the FSresidual
was calculated using the residual undrained or remoulded strengths), and to the post liquefaction condition of Lower
San Fernando dam, where FSresidual was based on the residual undrained strength of the liquefied sandy soils. Most
cases were best predicted by the rapid model.
5.1.3
Numerical Analysis
For slopes which do not displace or deform a lot on shearing, a reasonable guide to the displacements can be
obtained using numerical analyses. Some programs, eg. FLAC, are able to remain computationally stable at fairly large
strains. If nothing else, properly done, such analyses give a guide to the overall failure mechanics.
5.2
Post Failure Behaviour of Flow Slides and Flows
5.2.1 Empirical methods using travel angle.
Heim (1932) defined "fahrbschung " as the slope of a line connecting the crest of the source area with the toe of
deposits, measured on a straightened profile of the path (figure 67):

Source

Deposit

Figure 67 : A schematic path profile of a rock avalanche, showing the fahrbschung angle ()
Heim noted that, should one analyse the motion of a sliding block with constant frictional resistance from one end of
the profile to the other, the frictional coefficient would theoretically equal tan. Fahrbschung therefore is sometimes
called the "equivalent friction angle" of the slide. This is inaccurate, since the energy expenditure is characterized by
the displacement of the centre of gravity of the slide, not by the fahrbschung angle. Nevertheless, should the slide
motion be controlled by frictional forces only, with no pore pressure, the fahrbschung angle should be approximately
equal to the mean friction angle.
As noted by Heim and documented by Scheidegger (1973), Abele (1974) and others, the fahrbschung of large rock
avalanches is much less than the typical dynamic friction angle of dry broken rock (30 to 35). Further, the angle

  

apparently decreases with increasing magnitude of the event, as shown in figure 68. Rock avalanches from volcanic
sources were shown to be more mobile than others (Ui, 1993). The trend was first documented for rock avalanches
exceeding 1 million m3. However, it applies to certain mobile landslides which are much smaller, as shown for
example by Hutchinson (1995) for involving porous materials, such as chalk, by Evans and Clague (1988) for rock
avalanches in glaciated volcanic areas and Hungr and Kent (1995) for flow slides in mining waste. While the trend is
clear, its scatter is too large to permit reliable use for any but the most preliminary predictions of the travel distance.
Abele (1974) compiled a very detailed data base of the behaviour of rock avalanche from the Alps and attempted to
sort the volume fahrbschung relationship with respect to the broad geological domain and he shape of the deposition
area. Further sorting of the fahrbschung - volume data based on path morphology was presented by Nicoletti and
Sorriso-Valvo (1991). Such sorting reduces the scatter of the correlation to some degree.











  





Figure 68 : The relationship between the tangent of the fahrbschung angle


and volume for rock avalanches (Scheidegger, 1973)
Abele's data was used by Li-Tianchi (1983) to derive correlation equations and confidence limits suitable for
predictions. Howard (1973) plotted H/L versus the total potential energy of the slide, finding a trend similar to
Scheidegger's in both the sense and scatter. Hungr (1981) repeated Scheidegger's plot, expressing " fahrbschung " as
the vertical angle between the centres of gravity of the source and deposit. The trend is similar and the scatter is
undiminished. In addition, estimation of the position of the gravity centres is difficult in most cases.

   
    5), who argued that rock avalanches flow as a fluid,
depositing on the flat segment of the path, as shown in figure 67. If a frictional model was to apply, a dry rock
avalanche deposit would theoretically remain within a 32 vertical angle of the debris, corresponding to the dynamic

   ! !"     #
  
  
     $
 
"excessive travel distance" Le. This can be shown to increase with increasing volume of the event. Again, the trend is
evident, but the scatter of the data reduces the practical usefulness of this approach. The approach is particularly
inaccurate in cases where the slope profile is gradual, not bi-linear as in figure 67.
Corominas (1996) also applied this concept. He gathered a database of 204 landslides, and classified these into four
classes rockfalls and rockfall avalanches; translational slides; earth flows, mud flows and mud slides; and debris
flows, debris slides and debris avalanches; and took account of whether the slides were obstructed or channelised in the
travel path. Unlike other authors, the database included most slides smaller than 100,000 cubic metres ie. in the normal
engineering range. He found that the trend for the reduction of travel angle a (which is equivalent to the
 %
with increasing volume exists for all classes of slide. However, there is very substantial scatter in the data, which
cannot be substantially reduced based on morphological parameters only.
Corominas (1996) developed equations, with confidence limits for each class of landlside and degree of obstruction.
Table 15 summarises his data which applies to the equation
Log (tana) = A + BlogV
Where

a
V

=
=

(26)

travel angle
landslide volume in m3

Corominas (1996) includes data on translational slides which remain intact. This shows the apparently rather
illogical trend that only large landslides may have small travel angle, and ignores the pre and post failure strength, or
factor of safety, on the surface of rupture, and the authors believe the method is not applicable to such cases.

Wong and Ho (1996) and Finlay et al (1999) present data on small landslides in Hong Kong. Finlay et al (1999)
include data for cuts, fills and retaining wall failures which is summarised in figure 69.
It can be seen that the range of travel angles is very large, particularly for small slides. This reflects the different
mechanisms and degrees of obstruction. The very low travel angles for some fills may reflect the cases where collapse
has occurred on shearing of loose fills, but there are also cases of wash-out (where surface water has flowed over the
fill). The high values of F (high travel angle) represent cases of dilatant failure, or mostly falls from very steep cuts.

TABLE 15 : PARAMETERS FOR ESTIMATING THE TANGENT OF TRAVEL ANGLE A FOR


DIFFERENT CLASSES OF LANDSLIDE (COROMINAS 1996)
Landslide Type
All landslides
Rockfalls
Rockfalls
Debris flows
Debris flows
Earth flows
Earth flows

Path Type

All
All
Unobstructed
All
Unobstructed
All
All Unobstructed

-0.047
0.210
0.167
-0.012
-0.031
-0.214
-0.220

-0.085
-0.109
-0.119
-0.105
-0.102
-0.070
-0.138

Standard Deviation
Of log (tana)
0.161
0.123
0.073
0.137
0.093
0.131
0.074

These methods provide a valuable first pass estimate of what may happen and can be applied with some
judgement. For example if the fill is known to be dense and dilatant, larger travel angles will be applicable, than if the
fill is loose and contractant.

Figure 69 : F = tan (travel angle) = tana for landslides which break-up on sliding (Finlay et al 1999)
5.2.2
Correlations based on area/volume relationship
With the assumption that rock avalanche deposits of various sizes are geometrically similar, the volume (V), and
area (A), of the deposit should be related by the equation:
A = cV2/3

(27)

where c is a constant.
Equation 27 is plotted by a dashed line in figure 70, which includes data points for 40 rock avalanches. Apart from

a considerable random scatter, the general trend of the data fits equation 25 with a c = 12.0.
The random scatter probably results from topographic complexities such as channeling, flow obstructions and
branching, but also from the influence of the character of the material covering the flow path. For example, all four of
the points showing the highest degree of spreading crossed glaciers in their paths.
Another source of random scatter is inaccuracy in the estimates of volume. Only a very small percentage of rock
avalanche deposits has been examined in sufficient detail to provide accurate estimates of deposit thickness and it is not
uncommon to find differences of a factor of two or more between estimates by different authors.
The geometrical similarity trend leads directly to the prediction of the excess travel distance as a function of volume,
if it is assumed that rock avalanche profiles are typically bi-linear (Davies, 1981, Hungr, 1990). Hs's "Excessive
Travel Distance" is defined as:

Le = L

H
tan 32

(28)

If we assume that 32 is the average angle of the source/travel segment of the path, then Le must simply equal the length
of the deposit, measured in the direction of motion. This can be expressed as a function of the deposit area A and mean
width B or the mean aspect ratio R (length over mean width):
Le

A
=
B

RA (29)

(29)

Where:
R = Le/B

(30)

Le = ( RcV 2 / 3 = 3.47 R 1/ 2V 1/ 3

(31)

Substituting for A from (25) (c = 12.0 from figure 70)

DEPOSIT AREA (1,000m2)

10000

1000

100

10
100

1000

10000

100000
3

DEPOSIT VOLUME (1,000m )

Figure 70 : A plot of deposit area against volume for 35 rock avalanches from North America and Europe.
The dashed line has a slope of 2/3
5.2.3
Sliding block (lumped mass) models
The sliding block model, mentioned in section 5.1 above, can also be used for an approximate analysis of flowing
landslides. This approach has long been in use by snow avalanche workers (e.g. Perla et al., 1980).
Koerner (1976) adapted the resistance equations developed by Voellmy (1955) for snow avalanches. The movement

resistance of the block is assumed as composed of a frictional term and a turbulence term, dependent on the square of
velocity and inversely dependent on the flow depth (this term is a measure of kinetic energy per unit volume). The
resisting longitudinal force, T, acting on the base of a block is given as:

T = Mg cos tan +

v2

(32)

Where M is mas    


      the friction angle (including any pore- 
   
          

   !   "


  
model offers a good simulation of velocities for rock avalanches and debris flows. It is possible that the second term in
equation (32) represents rate dependent undrained shear strength of a rapidly shearing zone at the base of a flow (Ayotte
and Hungr, 2000).
Since the flow resistance depends on two parameters, an infinite number of pairs of the coefficients can be used to
obtain a given total displacement. However, each pair produces a different velocity profile (figure 71). The fastest
velocities are produced when the frictional term is dominant, while slides controlled by turbulent resistance move over
long paths with limited velocities. An appropriate choice of a unique pair can only be made if there are some
independent velocity observations along the path. The Koerner approach was used by the Garibaldi Review Panel to
estimate the runout of a rockslide from the Barrier in southern British Columbia in Canada (Hardy et al., 1978).

Figure 71 : A dynamic analysis of a rock avalanche using a two-parameter lumped mass

(sliding block) model (MacLellan and Kaiser, 1984).


McLellan and Kaiser (1984) attempted to find a method for assigning a pair of resistance coefficients "a priori", to
determine the total runout in a given path. They found the most consistent results with a friction coefficient equal to the
tangent of the mean runout slope.
Sassa (1988) defined a one parameter model, where the "apparent friction coefficient" is a function of a
pore-pressure coefficient, B. This coefficient is assumed to describe the increase in pore-pressure of the partly saturated
soils lying in the path of the avalanche as a result of over-riding and rapid loading by the flow front. Sassa suggests a
laboratory testing procedure for determining B. An alternative approach is to consider the B-coefficient as an empirical

factor, to be obtained by back-calculation. The form of variation of B along the path is a function of the degree of
saturation of the path soils.
Hutchinson (1986) and DeMatos (1986) independently derived models using a frictional term, coupled with a
one-dimensional consolidation algorithm to predict the gradual dissipation of excess pore pressure in an initially
liquefied basal layer. The resulting resistance equation is again a two parameter scheme similar to the Koerner model.
The velocities derived by DeMatos (1986) for certain case histories appear unrealistically high.
The lumped-mass approach cannot account for lateral confinement and spreading of the flow and the resulting
changes in flow depth. It should therefore be regarded as an empirical approach, suitable only for comparing paths
which are very similar in terms of geometry and material properties.
5.2.4
Continuum mechanics models.
Continuum mechanics models of rapid landslides use techniques developed for analysis of the flow of fluids in open
channels. There are, however, important differences between fluids and other, earth materials, even if the latter are
saturated and highly disturbed:
a) Soils and rocks are heterogeneous and discontinuous. Even when highly disturbed, they may have widely
different properties in various regions. It is not uncommon to find a zone of fluid-like liquefied soil adjacent to
a coherent block of essentially dry soil or rock.
b) Landslide materials have complex rheologies, which often change in time or with displacement due to
drainage, mixing, comminution or particle reorientation..
c) Landslides entrain and often discard material during motion.
In addition to the above, landslide paths are often much steeper and more varied than channels considered in most
hydraulic calculations and landslide motion is highly unsteady.
These characteristics make the analysis of landslide motion exceedingly complex. In general, landslide dynamic
models cannot claim to account for the detailed structure of the flowing mass. Instead, it is necessary to find an
equivalent fluid whose rheological properties are such that the bulk behaviour of the flowing body simulates the
expected bulk behaviour of the prototype landslide. The properties of the equivalent fluid do not correspond to those of
any of the slide components. For example, a flow slide consisting of essentially dry sand moving on a thin basal layer
of liquefied soil may be simulated using the Bingham rheology. However, the Bingham properties of the equivalent
fluid are neither those of the dry sand (which is frictional) nor those of the liquefied basal layer.
Most presently available solutions applicable to landslides use a form of the governing equations which results from
integrating the dynamic equilibrium and continuity relationships in the vertical or normal direction (perpendicular to the
flow bed), as shown in figure 72. The resulting St Venant equations relate to an elementary slice for a two-dimensional
problem or column for a three-dimensional one.

Figure 72 : The basic element used to derive an integrated Lagrangian solution to the St. Venant flow
equations (Hungr, 1995)
The Eulerian approach, widely used in hydraulics, references the calculations to a reference frame fixed in space.
The general St.Venant equation for the dynamic equilibrium of a two-dimensional element can be derived from the
principles of momentum conservation:

dv dv
T
dH
+
= g sin
gk cos
dx dt
dx
H

(33)

here, v is the mean flow velocity within the reference column, x is the tangential distance from a fixed origin, is the
bed slope angle, H is the flow depth (measured normal to the flow bed), is the flow density, T is the shear force acting
between the bed and the base of the column, depending on the assumed rheology as described in Section 5.2.5 below
and is a correction coefficient to account for the shape of the vertical velocity distribution.
The coefficient k is defined here as a ratio between the tangential and normal average total stress within the column.
It is normally assumed to equal 1 in fluids (hydrostatic stress condition) and therefore k does not appear in the
St.Venant equation when used in hydraulics. However, the stress state in flowing earth materials may not be
hydrostatic as discussed below.
Equation (33) is a simple restatement of the momentum conservation principle. The two terms on the left side are
the convective and local acceleration. The first term on the right is the tangential component of gravity, the second is
the friction slope and the third is the resultant of the pressure gradient corresponding to the local slope of the free
surface.
For continuity, any volume change in the reference column must equal the net inflow or outflow of material from
the element.
Given the highly unsteady nature of landslide motion, it is advantageous to use a moving Lagrangian reference
framework (Potter, 1972, Savage and Hutter 1986). The distance x in this case is measured relative to a moving origin,
attached to the centre of gravity of each column. This eliminates the convective acceleration term in equation (33) to
yield:

dv
T
dH
= g sin
gk cos
H
dx
dt

(34)

The Lagrangian form of the St.Venant equation is simple to integrate numerically. It also facilitates the use of
different material properties in different parts of the flow to simulate heterogeneity. It also makes it possible to account
for entrainment/deposition of material and to keep track of internal strain in the sliding body which determines the
magnitude of the longitudinal pressure coefficient k (Hungr, 1995).
Three-dimensional equivalents of equations (33) and (34) can be derived using the same principles.
5.2.5
Rheological Constitutive Relationships
The resisting force, T, in equations (33) and (34) is a function of mean velocity, flow density and depth, derived by
an integration of the rheological constitutive relationship of the eqivalent fluid. Several different rheologies have been
considered in the literature on landslide dynamics:
Plastic rheology: Earth liquefaction studies often use the concept of steady state strength, defined as a material
property. According to this, the resisting T is a constant, uninfluenced by neither velocity nor flow depth. The velocity
profile is constant, as the flows shears at the base where the applied shear stress is largest.
Frictional rheology: The resisting shear forces at the base of the flowing mass are assumed to depend on the
effective normal stress, but not on velocity. This assumption implies that the vertical velocity profile is indeterminate.
In general, however, it is expected that shear strains will be concentrated in a narrow zone at the base of the flow, where
the material is finest and where saturation and pore-pressure may exist.
The frictional model can be characterized by one parameter, the "bulk friction angle" b , which is related to the
effective dynamic friction angle ' through the use of a pore pressure ratio, ru (Cruden 1980, Sassa, 1988):
b = arctan [tan ' (1-ru)]

(35)

The effective dynamic friction angle of loose granular debris equals approximately 32. Full saturation (without
excess pore-pressure) would produce a bulk friction angle of about 17. Lower values of the angle imply the existence
of partial liquefaction at the base of the slide mass.
Turbulent flow: Turbulent flow is the principal model for analysis of fluid flow in rivers and channels, based on
Mannings equation. The resisting stress is a function of mean velocity squared and one-third power of the inverse of
the flow depth. This model probably applies only to the most dilute forms of debris flow.
Bingham Model: The resisting shear stress is assumed to depend on a constant strength and a viscous term
dependent on the velocity and the inverse of the debris sheet thickness. There are two material constants, a yield shear
strength and a Bingham viscosity. The flow profile includes a rigid plug, underlain by a parabolic velocity distribution

of laminar viscous flow. With zero shear strength, the flow becomes Newtonian. With zero viscosity, it becomes
plastic.
Voellmy rheology: This two-parameter model was developed by Voellmy (1955) for use in lumped-mass modelling
as described above, but can be used with good success in more detailed models of debris flows (Rickenmann and Koch,
1997), rock avalanches (Hungr and Evans, 1996) and some flow slides (Hungr et al., 2000).
Models with several parameters: Some authors have proposed models with several terms, incorporating plastic,
frictional, dilatant, viscous and turbulent effects (OBrien et al., 1993, Norem et al., 1990).
Any given total displacement of a landslide can be simulated by choosing the appropriate values of the resistance
coefficients for each model. However, the behaviour of the landslide, e.g. its velocity, degree of longitudinal spreading
and distribution of deposits along the path, will be different for each rheology (Ayotte and Hungr, 2000). The selection
of the appropriate rheology should be based on back-analysis of actual case histories. Constitutive relationships with
fewer parameters are to be preferred wherever possible, as they can be more easily calibrated.
There are several special factors which can be neglected in routine fluid flow calculations, but should be considered
during analysis of rapid flow slides.
Many landslides move on thin basal layers of weak material, while the bulk of the slide mass is relatively strong and
rigid. The slide mass is therefore capable of sustaining non-hydrostatic stress states such as those which exist in solid
or plastic media. Should the sliding body be assumed frictional and plastic, the coefficient k in equations (33) and (34)
may vary between the passive and active Rankine state limits, i.e. possibly in a range as wide as 0.2 to 5, depending on
the magnitude and sense of longitudinal strain acting within a column. The Lagrangian model allows one to keep track
of internal strain and to adjust the value of k accordingly (Hungr, 1995).
Centrifugal forces due to sharp changes of slope angle of the path sometimes have quite a strong influence on the
flow of materials with frictional behaviour and should be accounted for.
Gradual entrainment of material from the path of the flow can also be important. For example, moderately sized
rockfalls often entrain large volumes of talus, mobilized by impact liquefaction. A method of incorporating material
entrainment or deposition into the analysis was proposed by Hungr (1995). Material entrainment affects the flow in
different ways with different rheologies. In case of the frictional model, entrainment reduces the runout distance. In
case of models incorporating velocity dependent terms, entrainment may increase the runout distance (Hungr and
Evans, 1997). Entrainment of a saturated material also often changes the rheology of the flowing mass (Sassa, 1988).
Debris flows passing over saturated sediments can create a pressure wave in these sediments, which runs ahead of the
flow itself, and can fluidify the sediments. Another mode of increasing the mobility of mass movement, is air
entrainment or aerodynamic uplift. Some large landslides became airborn eg. Elm, Huascaran.
5.2.6
Some Examples of Dynamic Analysis
Perhaps the most well developed group of models dealing with the flow of earth materials is based on the Bingham
rheology. Jeyapalan (1981) concentrated on the special case of unsteady flow from a sudden breach of a dam (the dam
break problem) and derived solutions in terms of linearly viscous (Newtonian) laminar flow and visco-plastic
(Bingham) flow. A Bingham flow solution for routing of mudflow flood hydrographs in channels of limited slope was
formulated also by the US Corps of Engineers (1985). Fread (1988) presented a similar solution in terms of a power
law.
Dent (1982) modified an existing Newtonian unsteady flow program to accept a bi-linear rheology similar to the
Bingham model. His program (called BVSMAC) was used by Soussa and Voight (1991) to analyse several case
histories of rock avalanches.
The frictional rheology, where the resisting basal stress depends only on the normal stress, have long been favoured
in lumped mass models (e.g. Heim, 1932). A three-dimensional frictional model was developed by Sassa (1988), who
also introduced an undrained pore-pressure ratio.
Savage and Hutter (1986) used a two-dimensional Lagrangian frictional model to simulate the flow of dry sand.
Their algorithm is capable of simulating non-hydrostatic internal stress states appropriate for the flow of a granular
material, as opposed to a fluid.
Norem et al. (1990) present a two-dimensional model for submarine flow slides based on a complex rheological
formula combining frictional and viscous effects. This has also been used to model rock avalanches (J. Locat, pers.
com.).
It appears that dynamic analysis of debris flows will require a different approach from that applicable to other types
of flows. Iverson (1997) used detailed large-scale physical experiments, to show that longitudinal heterogeneity plays
an important role in controlling the dynamic behaviour of debris flow surges. He was successful in simulating an
experimental surge in a flume, using a non-homogeneous Lagrangian model. A boulder front, represented by a wedge
of frictional material, is followed by more dilute flow, represented in this case as an inviscid fluid. Hungr (2000)
analysed a similar flow regime in the steady state, using the theory of uniformly progressive flow.
Due to the different rheological assumptions, it has not been possible to compare one model with another. Hungr
(1995) developed an universal model, "DAN", which can use a variety of rheologic kernel formulas. Seven alternative
rheological functions can be used with the present version of DAN: plastic flow, friction flow, Newtonian laminar

flow, turbulent flow, Bingham flow, Coulomb viscous flow and frictional / turbulent "Voellmy" fluid. The model
can also account for lateral confinement of the path, non-hydrostatic intermal stress and for gradual
entrainment/deposition of material during motion. It is also possible to account for changing rheology along the path,
or longitudinally within the flowing mass. Calibration of this model for a variety of flow-like landslide types is in
progress (Ayotte and Hungr, 2000).
Figure 73 gives an example of the results of analysis of the Mt. Cayley rock avalanche using DAN.
In order to achieve practical predictive capabilities, future work in the field of landslide dynamics must concentrate
on controlled and systematic back-analysis of actual case histories.
A

E LE V A T IO N (m )

1000
W ID TH

200

800
600

100

400
E R O S IO N

P A T H W ID T H (m )

300

1200

200
0

1000

2000

D IS T A N C E (m )
P rofiles at 20 sec. intervals, depth exaggerate d 10x

V E LO C IT Y (m /s)

40
F ront

30

T ail

20
10
0
0

1000

2000

A V G . D E P O S IT
T H IC K N E S S (m )

D IS T A N C E (m )

10
8
6
4
2
0
0

1000

2000

D IS T A N C E (m )

Figure 73 : An example of a dynamic analysis of a rock avalanche involving pyroclastic breccia at Mt.
Cayley, a Quaternary volcano in British Columbia (Evans et al., 2000). (A) flow profiles plotted at 20
second intervals (note erosion of material from the path). (B) Velocity profiles. (C) Deposit distribution.
5.3`
Recommended Approach to Estimating Travel Distance and Velocity for Engineering Scale Slopes
Figures 74 and 75 show several situations where calculations of post failure deformation, travel distance and
velocity may be required. The following methods are suggested to do this:
(a) Cut slope, or fill, on horizontal ground (figure 74(a))
Use conservation of energy method Khalili et al (1996) to estimate x. Judge y from x and the nature of the slide
material. For slopes which will break-up, use Corominas (1996) or Finlay et al., (1999). For more important slopes,
use numerical analysis.
(b) Cut, with steep section at the base (figure 74(b))
Use conservation of energy method to estimate x, but note that if the movement is large (FS residual <<1.0) then the
method will under-estimate x because it does not allow for depletion as the slide mass falls. Estimate a from

Corominas (1996) or Finlay et al., (1999) allowing for the uncertainty in the outcome for more important structures, use
dynamic analyses. A number of dynamic models of varying complexity exist, as reviewed in Section 5.2.
Unfortunately, none of these models has so far been thoroughly calibrated to permit its use in routine predictive work.
The only feasible approach at present is to select a model and calibrate it locally, using several prototype landslides
similar to the event under investigation (eg. Hungr et al., 1999). Once calibrated, the model can then be used for
prediction of runout distances, depths and velocities. At present, such a process still has the nature of an applied
research project, rather than routine analysis.
(c) Cut, with steep, planar potential failure surface (figure 74(c)).
Check 'residual on failure surface versus fs. If fs > 'residual then assume all slide material falls to the base. Estimate
a from Corominas (1996) or Finlay et al., (1999). For more important structures use dynamic model.
(d) Shallow translational slide on a natural slope (figure 75(d)).
Base estimate on performance of similar slopes in the area. If there is no data available, check 'residual or residual
undrained strength (su) versus fs. If fs > su (or fs > 'residual), assume flow will occur. Estimate a from Corominas
(1996).
X

P O S T S L ID E

CUT SLOPE

?
POST

(a)

S U R FAC E O F
R U P TU R E

S L ID E

CUT SLOPE
X

PO S T SL
ID E

(b)

CUT SLOPE

S U R FAC E O F
R U P TU R E

S L ID E D E B R IS

PO

ST
SL
ID
E

FS

CUT SLOPE

(c)

S L ID E D E B R IS

Figure 74 : Estimation of travel distance for cut slopes


(e) Fill slide on a natural slope (figure 75(e)).
Use conservation of energy method to estimate if movements will be small or large. If small, accept those figures.
If large, assume flow may occur, and use Corominas (1996) and Finlay et al., (1999) to estimate the range of a. If fill
is loose, adopt low a values. If well compacted, adopt medium values.

(f) Reactivated landslides (figure 75(f)).


Use long term monitoring, but if this is not available, use conservation of energy method, taking account of the
possible reduction in factor of safety from initiation of sliding ie. any further strain weakening due to re-shearing on
rupture surface, internal deformations, shearing of lateral or 3 dimensional constraints. In most cases there will be no
reduction in strength (which will be 'R at least not calculable. Some movements, particularly retrogressive sliding, may
still occur. For more important slopes, use numerical analysis.
Above all, use common sense and observation of what has happened in similar slopes in the area. Accept that the
answers will be imprecise, and allow for the uncertainty in your analysis and decision making.

a
POS T S LIDE

(d)

NATU RAL S LOPE

FS

SLIDE DEBRIS

NATU RAL S LOPE

FILL

(e)

SLIDE DEBRIS
POS T S LIDE

REA CTIVATING
LA NDSLIDE

(f)

Figure 75 : Estimation of travel distance for natural slope failures and fills
5.4

Dynamic Analysis of Rock Fall


Rock fall, or fragmental rock fall (Evans and Hungr, 1993) is characterized by independent motion of individual
rock fragments, represented as rigid particles rolling and bouncing over the slope surface. Maximum runout of frequent
rock fall can generally be identified easily as the distal limit of talus deposits. It is more difficult to estimate the runout
of rare, large diameter particles, which tend to roll beyond talus margins. An empirical method for this purpose was
suggested originally by Lied (1977) and confirmed by Evans and Hungr (1993) as shown in figure 76. This relies on
projecting a vertical angle from the apex of the talus accumulation. The angle is of the order of 27 on coarse active
talus cones, but may reach as low as about 24 on snow or grassy slopes.
Dynamic analysis of rock fall using the principles of ballistics and collision theory was introduced by D.R. Piteau
(Piteau and Peckover, 1978) and subsequently developed by a large number of workers. All solutions use simple
ballistic parabola while the fragment is in the air. The various algorithms differ in the way they represent the interaction
between the fragment and the ground surface.
An exact solution of a collision of a spherical particle with a plane substrate was derived by Goldsmith (1950) and is
summarized in a slightly modified form in figure 77. The particle approaches with an incident linear velocity V,
consisting of a tangential component Vt and a normal component Vn, rotating at an angular velocity . After the

collision with the surface, all three components are modified to their rebound equivalents, Vt, Vn and . Some
energy is lost. The transformation matrix between the incident and rebound velocity tensor depends on whether
  

  

   #   "     $   % &  

 ' 

angle is less than the friction angle, , no slip occurs and the tangential and rotational momenta are partitioned using
matrix B. If slip occurs, additional energy is lost at the point of contact and the collision is described by Matrix C. In
both cases, the energy loss in the normal direction is controlled by a normal restitution coefficient en. A. Mellal
(unpublished research at the University of British Columbia) added a tangential restitution coefficient, to account for
energy losses other than tangential slippage, such as crater formation or breakage of surface asperities.

Figure 76 : Profiles of surveyed rock fall paths from British columbia, plotted with the apex of the talus cone
as a common fulcrum. Note the constant rock fall shadow angle, 27.5o, indicated by the dashed line
(Evans and Hungr, 1993).

tan =

if Cf > tan (limited sliding)

I (Vt R )
( I + mR 2 )(1 + e nV n )

[A]

0
0

V n' e n
V n

2
mR
IR
'
V
Vt = 0
et
2
2 t

I + mR I + mR

'

mR
I

et
0
2
2
I + mR I + mR

[B]

if Cf < tan (unlimited sliding)


en
0 0

V n'

+
et 0 V n
Cf (1 e n )

Vt' =
Vt
mR
Cf (1 + e n )
0 1

I
'

[C]

R: radius of the block


I: inertia momentum of the block
m: mass of the block
en: normal restitution coefficient (ratio of normal velocities after/before impact)
et: tangential restitution coefficient (ratio of tangential velocities after/before impact)
Vn' , Vt' , ' : normal, tangential and rotational velocities after impact
V n , Vt , : normal, tangential and rotational velocities before impact
Cf = tan : contact friction coefficient

Figure 77 : Idealised model of the collision of a fragment with the substrate (after Goldsmith, 1950)
The classical solution assumes that the restitution coefficients are constant, implying that the collision energy loss is
of a viscous character. In rocks, however, plastic deformation occurs as a result of cratering, crushing and breakage.
The restitution factors should therefore be scaled as function of the normal momentum of each impact.
The largest group of existing algorithms for rock fall ballistics, the lumped mass solutions, derive from
simplifications of the theory summarized in figure 77. The earliest model by Piteau and Peckover (1978) neglected the
rotational velocity and used only one constant restitution coefficient. A random element was added, however, by
perturbing the slope angle at the point of contact randomly through a set range, to account for roughness. A similar
roughness function is included in the model developed by the Colorado Department of Transport, CRSP (Pfeiffer and
Bowen, 1989), which also includes rotational velocity and two restitution coefficients, scaled to the incidental velocities
through empirical functions. This model has received a fair amount of calibration, but is very sensitive to the selected
roughness value.
Evans and Hungr (1993) neglected both roughness and rotational velocity, but introduced two separate random
restitution coefficients. If such coefficients are allowed to vary randomly, through a wide enough range, they may
account for the entire phenomenon approximately. The normal restitution coefficient should range to values greater
than 1 in such a model, as the rotational momentum may be transformed into linear in some impacts. The same authors
incorporated a plastic deformation model due originally to Falcetta (1985). This facilitates sustained reduction of the
restitution coefficients once the impact momentum exceeds a certain critical value. The effects are shown in figure 78.
Figure 78(a) shows a part of a trajectory of a particle falling from a large cliff and bouncing over the surface of talus,
with constant restitution coefficients. The great rebound after the first impact is unrealistic, as experiments showed that
as much as 75-86% of energy is consumed in such a major impact. The subsequent lighter impact consume much less
energy. The behaviour of the fragment is much more realistic in the second example (figure. 78(b)), run using the
Falcetta (1985) plastic deformation model.
A number of variations of this kind of modelling have appeared in the literature recently. Some three-dimensional
models exist, which attempt to simulate the lateral dispersion of rock fall paths (e.g. Descoudres and Zimmermann,
1987). All of the complexities of collision behaviour described above apply in equal measure in the lateral dimension.
All lumped mass models ignore the shape of the particles, which can only be incorporated indirectly through
random variation of the controlling factors. Several more sophisticated rigorous models now exist, which simulate
the motion of actual shapes (e.g. Bozzolo et al., 1988). One of the most advanced models was developed by the BRGM
in Marseille, France (Leroi, 1997). The model (figure 79) provides an exact solution to the movement of a three-

dimensional shape, including plastic deformation at the point of impact. These models are very demanding
computationally and must still accept a large degree of idealization, particularly in representing the surface roughness
of the slope.
The most important task facing rock fall researchers now is to provide wide-ranging, comprehensive calibration of
the models, against full scale field observations, including the full range of slope surface types.
















 




 


Figure 78 : Example of the effect of plastic deformation on rock fall fragment rebound. Both diagrams show
the trajectory of a fragment 2.6 tonnes in mass, dropped from a 100m cliff onto a talus surface. The normal
restitution coefficient is 0.7, the tangential 0.8. In (a), the collision is elastic. In (b) it incorporates the
Falcetta plastic model described by Evans and Hungr (1993).

Figure 79 : A rigorous three-dimensional model of a falling rock fragment developed by the


BRGM, Marseille, France (Leroi, 1997).
6.
6.1

MANAGEMENT OF SLOPE HAZARDS

Methods Available for Slope Safety Management


The methods which are available for managing the hazards from natural and constructed slopes include:
(a) Factor of safety
(b) Observational, usually combined with factor of safety
(c) Zonation and avoidance
(d) Risk management, qualitative or quantitative.
These are discussed below. We believe that where it is practicable to assess the consequences of sliding, qualitative
or quantitative risk assessment methods should be used routinely both for natural and constructed slopes. Even if factor
of safety approaches are used, estimates of travel distance and consequences of failure, if failure were to occur, should
be made, and these used to assist in selecting adequately conservative factors of safety.

6.2

Factor of safety approach


The factor of safety approach has been used for the majority of management of individual constructed slopes.
The method involves defining the geometry, shear strengths and pore pressures for the slope, calculating the factor
of safety, and comparing this to a required factor of safety. Usually analyses are carried out to assess the sensitivity of
the answer to the input assumptions. The required factor of safety is usually related, to the conservatism of the input
parameters, and in a general sense, to the consequences of failure. Table 16 gives the typical requirements for
embankment dams. These usually have as inputs, relatively conservative estimates of strengths and pore pressures.

TABLE 16 : TYPICAL LOADING CONDITIONS, REQUIRED FACTORS OF


SAFETY AND SHEAR STRENGTHS FOR EMBANKMENT DAMS
Loading condition
Steady state at reservoir full supply level
Rapid drawdown
Construction

Required factor
of safety
1.5
1.2 - 1.3
1.2 1.3

Shear strength for evaluation


Effective stress
Minimum composite effective stress and undrained
Undrained

Implicit in the factors of safety, is the likelihood of breaching of the dam given the factor of safety falls below 1.0.
Hence the drawdown factor of safety required is lower than for steady seepage, because the reservoir will be lower, and
a large freeboard would make overtopping of the dam unlikely even if instability occurs. For the construction loading,
the lower factor of safety is reflecting the lower consequences of failure, because there is no reservoir behind the dam.
For slopes other than dams, there are no generally accepted standards. Sllfors et al (1996) report on the
Recommendations for Slope Stability Analysis Royal Swedish Academy of Science Commission on Slope
Stability Report 3:95, 4:95, 5:95, Stockholm 1995. This presents a chart of required factors of safety which is
reproduced in Table 17.

TABLE 17 : CHART OF REQUIRED CALCULATED FACTORS OF SAFETY IN VARIOUS STEPS IN


AN INVESTIGATION WITH RESPECT TO LAND USE. (SLLFORS ET AL (1996), FROM ROYAL
SWEDISH ACADEMY OF SCIENCE, 1995).
Phase
New
Development

Use of land
Existing structures
Other land

Undeveloped
land

Geotechnical
inspection
and rough estimate

Detailed investigation
must be
carried out

Fc > 2 +
Fc > 1,5

Fc > 2 +
Fc > 1,5

Fc, Fcomb
and F > 1

Detailed
Investigation
And advanced
Investigation

Fc 1,7-1,5 +
Fcomb 1,45-1,35
F 1,3 (sand)

Fc 1,7-1,5 +
Fcomb 1,45-1,35
F 1,3 (sand)

Fc 1,6-1,4 +
Fcomb 1,4-1,3
F 1,3 (sand)

Fc, Fcomb
and F > 1
(If surrounding land
is unaffected)

Fc, Fcomb
In-depth
Fc 1,5-1,4 +
Fc 1,4-1,3 +
Fc 1,3-1,2* +
Investigation
and F > 1
Fcomb 1,35-1,30
Fcomb 1,30-1,20
Fcomb 1,2-1,5*
(and supple(If surrounding land
F 1,3 (sand)
F 1,3 (sand)
F 1,2-1,15 (sand)
mentary
is unaffected)
(combined with
*) lower values refer to
investigation)
restrictions)
existing structure
NOTE: For Fcomb, for each section of the slip surface the drained shear strength is compared to the undrained shear
strength and the lowest strength is used in the calculations.
The table has some useful concepts:

The required calculated factor of safety should be larger early in the investigations when the uncertainty in
parameters is great, reducing as the investigations reduce the uncertainties.

The required calculated factors of safety for soils where the undrained strength is controlling should be greater
than for frictional materials (reflecting the greater uncertanty and variability in the estimates of undrained
strength).

Lower factors of safety may be adopted if the consequences of failure are low (eg. for undeveloped land).

For many existing landslides where remedial works are to be carried out, it is common to seek an increase in the
factor of safety of say 0.25 or 0.3 (for normal engineered slopes) rather than an absolute factor of safety. This accounts
for the fact that many of the unkowns are already accounted for in such a moving slope. For large landslides, smaller
eg. 0.05 to 0.1 increases in overall factor of safety have been considered adequate. In these cases, often the factor of
safety for smaller slides at the toe of the slope is increased by a larger amount by the remedial works
Factor of safety approaches have the advantage that they are familiar, and properly applied, have resulted in low
incidence of failures. However they have the following limitations

The likelihood (probability) of failure is not defined and may be quite different for the same factor of safety in
different situations. For example Morgenstern and Ramly (2000) have demonstrated that for undrained
analyses of slopes in soft clays, a factor of safety of 1.5 may represent a probability of failure of about 10-2 to 2
x 10-2. For frictional materials, a factor of safety of 1.5 would usually represent a probability of failure of
about 10-4 to 10-5. The low probabilities of failure of dams, which are commonly designed for a factor of
safety of 1.5 for steady seepage, supports these latter figures (Foster et al 2000).

There is no systematic consideration of the consequences of failure ie no assessment of deformations and travel
distance, given the slope fails, and the potential for damage to property, and loss of life.
Most constructed slopes which fail do so because of unforeseen low strength zones, higher pore pressures than
anticipated, deficiencies during construction, or incorrect assumptions of the mechanics of the slope movement. These
deficiencies do occur, and slopes do fail, so it is unwise to not formally consider the consequences. Hence even if a
factor of safety approach is being used, the consequences, assuming failure occurs, should be assessed, to see whether
an appropriate degree of conservation is being used.
6.3

Observational Approach
The observational approach consists of monitoring of pore pressures and deformations, usually with reference to the
calculated factor of safety. For most slopes eg constructed cuts, fills, dams and embankments on soft clay, monitoring
is used as a back-up to factor of safety design ie as a check on the assumptions, rather than as a primary design tool.
For existing landslides, particularly large ones, it is often the main method of slope safety management. In these cases,
the slopes are monitored, and considered to pose an acceptable hazard, provided the movements remain in the primary
(or secondary) phases of creep movement (see Section 4.6). If movement approaches the tertiary creep stage, remedial
action is taken, or persons who are at risk, are evacuated. The observational approach has some limitations:
It is essential that the mechanics of the slope movement pre and post failure are understood. Situations where
rapid movement, possibly with large travel distances, may occur with little pre-failure deformation must be
recognised, and may not be suitable for this method of slope management. Examples of these are given in
section 2.1.
Measurements of movements must be taken at sufficiently small time intervals to be able to detect the on-set of
failure. There is no point relying on annual readings of movement of large natural slides as a hazard
management tool given the slide has a potentially brittle mechanism. (see Section 3.3.6). For these situations
daily or continuous monitoring may be needed.
The deformations may be too small to monitor sufficiently accurately, or the cost of doing so may be high. This
is particularly a problem for small, brittle slopes.
Where there is reliance on pore pressures as the guide to slope safety, it is essential the instruments are able to
monitor the transient conditions eg. water pressure on cracks, pore pressure rises due to rapid infiltration.
Again, it may be quite misleading to rely on readings taken daily or weekly, when the piezometric pressures
change rapidly (see Section 3.3.6).
6.4

Zonation and avoidance


For many natural landslides the hazard posed by the landslide is managed by zoning development to avoid the
hazard, or installing early warning systems so the population at risk can be evacuated from the hazard. This is
discussed in the Issues Lecture paperQuantitative Risk Assessment, by E. Leroi, K. Ho, W. Roberds in this
conference. For small natural landslides and resulting debris flows, etc, this is the only practical management system,
because it is impractical to calculate the factors of safety of each slope (and in any case, they will often be approaching
1.0 under certain circumstances eg. heavy rain, or snow melt), or to monitor the deformations of each slope.
6.5

Risk Management Approach


The risk management approach is summarized in figure 80.
The process formally requires considerations of the likelihood of sliding, the consequences, and hence the risk; and
comparison of there calculated risks to risk acceptance criteria, accounting for financial and loss of life issues.

SCOPE DEFINITION
RISK ANALYSIS

ESTABLISH BRIEF, PRO PO SED M ETHO DO LO G Y

HAZARD IDENTIFICATION
C LASSIFIC ATIO N O F LANDSLIDE e g slid e ,d e b ris flo w , ro c kfa ll
EXTENT O F LA NDSLIDE e g lo c a tio n, a re a , vo lum e
TRAVEL O F LA NDSLIDE
RATE O F M O VEM ENT e g c re e p , slo w , fa st

RISK ESTIMATION

CONSQUENCE ANALYSIS

FREQUENCY ANALYSIS

ELEMENTS AT RISK

ESTIM ATE FREQUENCY

PRO PERTY
RO ADS/ C O M M UNIC ATIO NS
SERVIC ES
PEO PLE
TEM PO RAL PRO BABILITY e g ve hic le s, p e rso ns

Q UA LITATIVE
Q UA NTITATIVE

VULNERABILITY
RELA TIVE DAM A G E
PRO BA BILITY O F INJURY / LO SS
O F LIFE

RELATE TO INITIATING EVENTS


RAINFALL
C O NSTRUC TIO N AC TIVITY
EARTHQ UAKE
SERVIES FAILURE/ M ALFUNC TIO N
AND/ O R HISTO RIC PERFO RM A NC E

RISK ESTIMATION
RISK = (LIKELIHOOD OF SLIDE) X (TEMPORAL PROBABILITY)
X (VULNERABILITY) X (ELEMENTS AT RISK)
CONSIDERED FOR ALL HAZARDS
RISK ASSESSMENT

RISK EVALUATION C O M PARE TO LEVELS O F TO LERABLE O R


AC C EPTA BLE RISK A SSESS, PRIO RITIES A ND O PTIO NS
C LIENT / O WNER / PO LIC Y M A KERS TO DEC IDE
TEC HNIC AL SPEC IALIST TO ADVISE

RISK CONTROL
RISK MANAGEMENT

TREATMENT OPTIONS
AC C EPT RISK
AVO ID RISK
REDUC E LIKELIHO O D
REDUC E C O NSEQ UENC ES
TRANSFER RISK

TREATMENT PLAN
DETAIL SELEC TED O PTIO NS

IMPLEMENT PLAN
PO LIC Y A ND PLA NNING

MONITOR AND REVIEW


RISK C HANG ES
M O RE INFO RM ATIO N
FURTHER STUDIES

Figure 80 : Risk Management Framework (Australian Geomechanics Society, 2000)

Figure 81 compares the factor of safety approach to a risk based approach for the case of a cutting behind a building.

CUT
SL O P E

(a)

BU ILD IN G

SM ALL SL ID E

M E D IU M S LID E

(b)

?
?

PE R SO N S

?
?

?
?

LA R G E SLID E

BU ILD IN G

Figure 81 : Comparison of a Factor of Safety Approach to a Risk Assessment Approach.


(a) Deterministic approach (b) Risk based approach
For the deterministic approach the questions are:
What is
- Geometry, Geology, Hydrogeology?
- Shear Strength?
- Pore Pressure?
What is factor of safety and senstivity to assumptions?
Is this acceptable? Taking into account the consequences?
For the risk based approach, the questions are:
What is
Slope Geometry, Geology, Hydrogeology?
Where? How big? And what is the probability of sliding?
What will the slide mechanism be?
How far will the slide travel? And how fast?
Will there be warning signs?
Will the slide reach the house/ How big? How fast?
Will there be persons in the house?
What is the vulnerability of the persons and the house?
What is the risk to life and property?
Is the risk acceptable?
The steps in the process will include
(a) Estimation of the likelihood and volume of sliding using (Leroi et al 2000, Mostyn and Fell, 1997):
historic frequency of occurrence
relation to geomorphology and geology
relationship of historic record of landsliding to rainfall or snow melt intensity and duration
relating the historic record of rainfall, slope geometry, geotechnical properties and landsliding

modelling piezometric levels, versus rainfall frequency, and calculating annual exceedance probability of
factor of safety below 1.0
formal probabilistic or reliability analysis in quantitative or qualitative terms.
(b) Calculating the deformation and travel distance of the landslide, and the velocities and depth of the slide over
the travel path, taking account of the mechanics of sliding and pre and post failure. When considering travel
distance, it is important to consider whether the slope could behave in a brittle manner, and experience rapid,
high velocity travel if it fails. This will be done as outlined in Section 5.
(c) Assess the elements at risk the property, buildings, and persons, allowing for temporal effects ie. the
proportion of time the persons are likely to be in the area affected by the landsliding (see Leroi et al 2000, Fell
and Hartford (1997)).
(d) Estimate the potential damage and loss of life (see Leroi et al 2000, Finlay, Mostyn and Fell (1999, 2000)).
(e) Calculate the risk for the individuals most at risk and to the population at risk using the equations
R(DI) = P(H) x P(S/H) x P(T/S) x V (L/T)

(36)

where
R(DI) is the risk (annual probability of loss of life to an individual)
P(H) is the annual probability of the hazardous event (the landslide)
P(S/H) is the probability of spatial impact ie. of the landslide impacting a building) given the event
P(T/S) is the probability of temporal impact (ie. of the building being occupied) given the spatial impact
V(L/T) is the vulnerability of the individual (probability of loss of life of the individual given impact).
For a case involving property damage the equivalent expression would be:
R(PD) = P(H) x P(S/H) x V(P/S) x E

(37)

where
R(PD) is the risk (annual loss property value)
P(H) is the annual probability of the hazardous event (the landslide)
P(S/H) is the probability of spatial impact (ie. of the landslide impacting the property)
V(P/S) is the vulnerability of the property (proportion of property value lost)
E is the element at risk (eg. the value of the property)
The procedure detailed above is becoming the one generally adopted.
(f) Assess the tolerability of the risk against the risk acceptance criteria.
Leroueil and Locat (1998) propose an approach for risk assessment and management which is similar to the one
described. They make use of the Geotechnical characterisation of slope movements a framework for organising
knowledge on slopes in which, in particular, they define pre-disposition factions, triggering factors, revealing factors,
and consequences of the movement or failure. In this context, the hazard is directly the probability of the triggering
factor to reach a critical value leading to failure; as for the elements at risk and their vulnerability, they should be found
in consequences of the movement.
There are a number of limitations to risk analysis and assessment for slopes and landslides (as spelled out by IUGS
Working Group on Landslides, 1997)
the judgement content of the inputs to any assessment may well result in values of assessed risks with considerable
inherent uncertainty
the variety of approaches that can reasonably be adopted to assess landslide risk can result in significant difference
in outcome if the same problem is considered separately by different practitioners
revisiting an assessment can lead to significant change due to increased data, a different method, or changing
circumstances
the inability to recognise a significant hazard and the consequential underestimation of the risk
the results of assessment are seldom verifiable, though peer review can be useful
the methodology is currently not widely accepted, and thus there sometimes is an aversion to its application
it is quite possible that the cost of the assessment may outweigh the benefit of the technique in making a decision,
especially where complex detailed sets of data are required
acceptable and tolerable risk criteria for slopes and landslides are not well established
it is difficult to accurately assess risk for low probability events.
It is considered by the authors that in many situations it will be sufficient to simply assess the likelihood of sliding in
very approximate terms, and assess the travel distance, elements at risk, potential damage and loss of life using simple
and empirical methods. The advantage of the risk management approach is that the process requires the formal
consideration of the question if the landslide did happen to occur, how far could it travel, how fast, and could it kill
people. Often, if the answer is yes, it may be that there are too great uncertainties in the understanding of the slope, or

too great a reliance on observational approaches, so more detailed investigations, or defensive measures to reduce the
hazard, on the consequences, are needed.
6.6

Hazard Reduction Measures


Hazard reduction measures include:
(a) Measures to improve the factor of safety (reduce the likelihood of sliding) including:
surface and subsurface drainage
slope modification
structural measures (eg. anchors, walls, piling).
(b) Measures to reduce the travel distance of the slide:

provision of non-strain weakening elements

catch walls and drains, catch dams.


(c) Warning systems
based on the triggering mechanism eg. rainfall duration and intensity
based on movement.
(d) Avoidance
These are discussed in Bromhead (1986), Schuster (995) Dunnicliff (1995), and Keaton and Beckwith (1996), Holtz
and Schuster (1996), Wyllie and Norrish (1996), Fell (1994), Leroueil and Locat (1998).
7.

RESEARCH AND DEVELOPMENT NEEDED

The authors see the following areas a having a priority in research and development:
Typology of landslides, to be used to constrain the applicability of various analytical techniques, depending on
the movement mechanism, and material characteristics of various landslide types.
Better understanding of the basic processes of creep in all its forms; and fatigue; and development of proper
modelling of the processes.
Development of practical methods for identifying contractant granular soils, and methods to analyse their
failure and post failure behaviour.
Development of a rigorous and logical approach to assessing the shear strength of fissured clays and weak
macro-fissured rocks such as shales and mudstones.
Development of an understanding of the pre failure deformation and other precursory signs of landslides of
different mechanisms and scale.
Detailed study of the reasons for brittle behaviour in large landslides, including study of the detailed mechanics
of the slides, including internal displacements, and the relationship between velocity of displacement,
mechanics, and shear strength on the surface of rupture.
Development of the understanding post failure movements, including reactivated slides, for different
mechanisms and scale of sliding, and progressive and retrogressive failure.
Development of practical methods, with a sound theoretical basis, for modelling infiltration of rainfall and
snow.
Development of commercially available computer programs, to calculate travel distance, velocity and energy
profiles based on available research programs, and calibration of these for a range of landslides. Further
refinement of the theoretical basis for these analyses, and methods for specifying what properties apply to a
range of landslides. Assembly of data-bases of post failure behaviour, to be used in an empirical manner, and to
calibrate the dynamic models.
Assembly of and development of methods and a theoretical basis for magnitude frequency relationships for
various types of landslides, using probabilistic methods.
Develop a broad framework for the behaviour of landslides in earthquakes, relating to their classification, the
topography, earthquake characteristics, and scale of landsliding.
The effect of salt, gypsum, or calcite on the strength of soil, and the effects of weathering on the shear strength
of clay.
There is an overwhelming need to integrate geology, geomorphology and geotechnical engineering in these studies.
Members of multidisciplinary research teams need to have a mix of theoretical and practical skills.
These groups need access to good quality case studies. Workshops, which bring together 20 or 30 active researchers
and practitioners, are a valuable way of exchanging ideas, and developing working linkages, and these should be
encouraged and fostered by the International Societies.
8.

ACKNOWLEDGEMENTS
The authors wish to thank the organisers of the GeoEng 2000 Conference for inviting us to prepare this paper.

9.

REFERENCES

Aas, G., Lacasse, S., Lunne, T. and Heg, K. (1986). Use of in-situ tests for foundation design in clay. Use of in-situ
Tests in Geotechnical Engineering, Geotech. Special. Publ. No.6, ASCE, pp. 1-30.
Abele, G. (1974). Bergsturze in den Alpen. Wissenschaftliche Alpenvereinshefte, No.25, Munich (in German).
Adachi, T., Liu, J., Koike, A. & Zhang, F. (1996). Finite element analysis of Biots consolidation in slope excavation
based on a constitutive model with strain softening. Proc. 7th Int. Symp. on Landslides, Trondheim. Editor K.
Sennesett, Balkema, Rotterdam, pp.1131-1136.
Alonso, E.E., Gens, A. Lloret, A. and Delahaye, C. (1995). Effects of rain infiltration on the stability of slopes. Proc.
1st Int. Conf. on Unsaturated Soils, Paris, Vol. 1, pp.241-248
Anderson, S.A. & Riemer, M.F. (1995). Collapse of saturated soil due to reduction in confinement. ASCE J. of
Geotechnical Engineering, Vol. 121(2), pp.216-220.
Anderson, S.A. & Sitar, N. (1995). Analysis of rainfall-induced debris flows. ASCE J. of Geotechnical Engineering,
Vol.121(7), pp. 544-552.
Australian Geomechanics Society (2000). Landslide risk management concepts and guidelines. Australian
Geomechanics Society Sub-committee on Landslide Risk Management. Australian Geomechanics, Vol. 35, No.1,
pp.49-92.
Ayotte, D., and Hungr, O. (1998). Runout analysis of Debris Flows and Debris Avalanches in Hong Kong. Report of
the Geotechnical Engineering Office, Hong Kong.
Azimi, C., Biarez, J., Desvarreux, P., Giuliani, Y. & Ricard, C. (1992). Mcanisme des glissements de terrains argileux
Bilan de surveillance sur plusieurs annes. 6th Intl. Symp. on Landslides, Christchurch. Editor D. Bell, Balkema,
pp.1903-1908.
Azimi, C., J. Biarez, P. Desvarreux, F. Keime, (1988). Prvision dboulement en terrain gypseux. Proc. Fifth Int.
Symp. on Landslides, Lausanne. Editor C. Bonnard, Balkema, Rotterdam, pp.531-536.
Azzouz, A.S., Baligh, M.M. and Ladd, C.C. (1983). Corrected field vane strength for embankment design. J.
Geotechnical Eng. ASCE., Vol. 109(5), pp.730-734.
Baltzer, A., (1875). ber die Bergstrze in den Alpen. Schweizerische Alpenclub, Bern, Jahrbuch 10: pp.409-456.
Banks, D.C. and Strohm, W.E. (1974). Calculations of rock slide velocities. Procs., Congress ISRM, Denver, 2,
pp.839-847.
Barton, N., (1982). Shear strength investigation for surface mining. 3rd Int. Conf. Stability in Surface Mining, C.O.
Brawner Editor, pp. 171-196.
Been, K. and Jeffries, M.G. (1985). A state parameter for sands. Geotechnique. Vol.35: 99-112.
Berntson, J.A. & Sllfors, G.B., (1984). Pore pressure variations in marine clay deposits. 4th Int. Symp. On
Landslides, Toronto, Vol. 1, pp.363-366.
Bertini, T., Cugusi, F., Delia, B. & Rossi-Doria, M. (1986). Lenti movimenti di versante nell Abruzzo Adriatico:
caratteri e criteri di stabilizzazione. 16th Convegno Natinale di Geotecnica, Bologna, Vol. 1, pp.91-100.
Binnie and Partners (1982). Fill slopes preventive works stage 2 studies, Report on tests on undisturbed samples of
compacted fill. Hong Kong Government (unpublished).
Bishop, A. W. (1955). The Use of the Slip Circle in the Stability Analysis of Slopes. Geotechnique, 5 (1),. Pp.7-17.
Bishop, A.W. & Bjerrum, L. (1960). The relevance of the triaxial test to the solution of stability problems. ASCE
Research Conference on Shear Strength of Cohesive Soils, Boulder, pp.437-501.
Bishop, A.W. (1967). Progressive failure - with special reference to the mechanism causing it. Proceedings of the
Geotechnical Conference on Shear Strength Properties of Natural Soils and Rocks, Oslo, Vol. Vol 2, pp.142-150.
Bishop, A.W. (1971). Shear strength parameters for undisturbed and remoulded soil specimens. Proceedings of the
Roscoe Memorial Symposium, Cambridge University. Edited by R.G. Parry; G.T. Foulis and Co. Ltd. Oxfordshire,
pp.3-58.
Bishop, A.W. (1973). The stability of spoil heaps. Quarterly Journal of Engineering Geology, Vol 6, pp. 335-376.
Bishop, A.W. and Bjerrum, L. (1960). The relevance of the triaxial test on the solution of stability problems. Proc.
ASCE, Research Conference on Shear Strength of Cohesive Soils. Boulder. Pp.437-501.
Bjerrum, L. (1967). Progressive failure in slopes of overconsolidated plastic clay and clay shales (Terzaghi
Lecture),.ASCE, Journal of the Soil Mechanics and Foundations Division, Vol 93 (No. SM5), pp.2-49.
Bjerrum, L. (1973). Problems of soil mechanics and construction of soft clays and structurally unstable soils
(collapsible, expansive and others) State of the Art report, Proc. 8th Int. Con. Soil Mechanics and Foundation
Engineering, Moscow. Vol.3, pp.111-160.
Blight, G.E., (1997). Destructive mudflows as a consequence of tailings dyke failures. Procs., Inst. Civil Engineers,
Geotechnical Engineering, 125: pp.9-18.
Blondeau, F. and Queyroi, D. (1976). Rupture de la tranche exprimentale de la Bosse-Galin (argile molle).
Bulletin de liaison des Laboratoires des Ponts et Chausses. Numro spcial III Stabilit des talus 2: Dblais et
remblais, pp.59-69.

Boudali, M., Leroueil, S. and Murthy, B.R.S., (1994) Viscous behaviour of natural soft clays. Proc. 13th ICSMFE,
New Delhi, Vol.1, pp.411-416.
Bozzolo, D., Pamini, R. and Hutter, K., (1988) Rockfall analysis a mathematical model and its test with field data.
In Procs., 5th. Int. Symposium on Landslides, Lausanne, Editor C. Bonnard. Balkema, Rotterdam, pp.555-560.
Bracegirdle, A., Vaughan, P.R. & Hight, D.W. 1992. Displacement prediction using rate effects on residual shear
strength. 6th Int. Symp. On Landslides, Christchurch. Editor D. Bell. Balkema, Rotterdam, pp.343-347.
Brand, E.W. (1995). Keynote paper: Slope instability in tropical areas. 6th Int. Symp. On Landslides, Christchurch,
Vol. 3, pp.2031-2051.
Brand, E.W., Premchitt, J. & Phillipson, H.B. (1984). Relationship between rainfall and landslides in Hong Kong.
Proc. 4th Int. Symp. On Landslides, Toronto, Vol. 1, pp.377-384.
Bromhead, E.N. (1988). The stability of slopes. Surrey University Press.
Bromhead, E.N. and Curtis, R.D. (1983). A comparison of alternative methods of measuring the residual strength of
London Clay. Ground Engineering, 16, pp.39-41.
Bromhead, E.N., Cooper, M.R. & Petley, D.J. (1998). The Selborne cutting slope stability experiment (CD-ROM The
Selborne data collection CD).
Brunsden, D., DoornKamp, J.C., Fookes, P.G., Jones, D.K.C. and Kelly, J.M.H. (1975). Large sale geomorphological
mapping and highway engineering design. Quarterly Journal of Engineering Geology, Vol.8, pp.227-253.
Bryant, J.M., Logan, T.C., Woodward, D.J. and Beetham, R.D. (1992). The use of seismic method in defining
landslide structure. Proc. Sixth Int. Symp. on Landslides, Christchurch, Editor D.H. Bell, Balkema, Rotterdam,
pp.33-40.
Burland, J.B., Longworth, T.I. and Moore, J.F.A. (1977). A study of ground movement and progressive failure caused
by a deep excavation in Oxford Clay. Geotechnique, Vol 27, (4), pp.557-591.
Cartier, G. & Pouget, P. (1988). Etude du comportement dun remblai construit sur un versant instable, le remblai de
Salldes (Puy-de-Dome). Laboratoire Central des Ponts et Chausses, Research Report No. 153, pp.153-130.
Cavounidis, S. (1987). On the Ratio of Factors of Safety in Slope Stability Analyses. Geotechnique, 37 (2), pp.207210.
Chan, D.H. and Morgenstern, N.R. (1987). Analysis of progressive deformation of the Edmonton Convention Centre
excavation. Canadian Geotech. J. 24, pp.430-440.
Chandler, R.J. (1984). Recent European experience of landslides in over-consolidated clays and soft rock. 4th
International Symposium on Landslides, Toronto, Vol. 1, pp.61-81.
Chandler, R.J. (1984a). Delayed failure and observed strengths of first time failures in stiff clays: a review.
Proceedings of the Fourth International Symposium on Landslides, Toronto, Vol. 2, pp.19-25.
Chen, W.F. (1975). Limit analysis and soil plasticity. Elsevier Scientific Publishing Co., New York.
Chigira, M. (1992). Long-term gravitational deformation of rocks by rock creep. Eng. Geol. 32, pp.157-184.
Chirapuntu, S. and Duncan, J.M. (1977). Cracking and progressive failure on embankment on soft clay foundation.
Proc. Int. Symp. On Soft Clay, Bangkok, Thailand, pp.453-470.
Christian, J.T., Ladd, C.C. and Baecher, G.B. (1992). Invited lecture: reliability and probability in stability analysis, in
Stability and Performance of Slopes and Embankments 11. ASCE Geotechnical Special Publication, No.31,
pp.1071-1111.
Clayton, C.R.I., Mattews, M.C. and Simons, N.E. (1995). Site Investigation, Blackwell Science Ltd., Oxford.
Collins, B. & Znidarcic, D. (1997). Triggering mechanisms of rainfall induced debris flows. Proc. 2nd Int. PanAmerican Symposium on Landslides, Rio de janeiro, Vol. 1, pp.277-286.
Collins, B. and Znidarcic, D. (1998). Slope stability issues of rainfall induced landslides. Proc. 11th DanubeEuropean Conf. on Soil Mech. And Geotech. Engng, Porec, Croatia, pp.791-798.
Cooper, B., Khalili, N. and Fell, R. (1997). Large deformations due to undrained strain weakening slope instability at
Hume Dam No. 1 embankment. Nineteenth International Congress on Large Dams, Florence, ICOLD. Vol. 2,
(Q73 R46), pp.797-818.
Cooper, M.R. (1988). A displacement based analysis of progressive failure by the Reserve Capacity method.
Proceedings of the Fifth International Symposium on Landslides, (Bonnard ed.) Lausanne, Switzerland, Balkema.
Vol. 1, pp.583-588.
Cooper, M.R. (1996). The progressive development of a failure slip surface in over-consolidated clay at Selborne,
UK. Proceedings of the Seventh International Symposium on Landslides, (Senneset ed.) Trondheim, Norway,
Balkema, Rotterdam. Vol. Vol 2, pp.683-688.
Cooper, M.R., Bromhead, E.N., Petley, D.J. & Grant, D.I. (1998). The Selborne cutting stability experiment,
Geotechnique, Vol. 48(1), pp.83-101.
Corominas, J. (1996). The angle of reach as a mobility index for small and large landslides. Canadian Geotechnical
Journal, 33, pp. 260-271.
Cruden, D.M. (1980). A large landslide on Mars: Discussion. Bulletin, Geol. Society of America, 1,91, p. 64.

Cruden, D.M. and Varnes, D.J. (1996). Landslide types and processes. In Landslides - Investigation and Mitigation,
Transportation Research Board Special Report No. 247 (A.T. Turner &. R.L. Schuster ed.), National Academy
Press, Washington DC, pp.36-75.
Cruden, D.M., (1991). A simple description of a landslide. Bulletin IAEG, No. 43, pp.27-29.
DElia, B., Picarelli, L., Leroueil, S. and Vaunat, J. (1998). Geotechnical characterisation of slope movements in
structurally complex clay soils and stiff jointed clays. Italian Geotechnical Journal, Anno XXXII, No.3: pp.5-32.
dOrsi, R., dAvila, C., Ortigao, J.A.R., Dias, A., Moraes, L. and Santos, M.D. (1997) Rio-Watch: the Rio de Janeiro
landslide watch system. Proc. 2nd Pan-American Symp. on Landslides, Rio de Janeiro, Vol.1: pp.21-30.
Davies, T.R.H. (1981). Spreading of rock avalanches by mechanical fluidization. Rock Mechanics, 15, pp. 9-24.
Davies, W.N. and Christie, H.D. (1996). The Coledale Mudslide, New South Wales, Australia A lesson for
Geotechnical Engineers. Proc. 7th Int. Symp. On Landslides. Editor K. Senneset. Trondheim, Balkema,
Rotterdam, pp.701-706.
Dawson, R.F., Morgenstern, N.R. and Stokes, A.W. (1998.) Liquefaction flowslides in Rocky Mountain coal mine
waste dumps. Can. Geotech. J., 35, pp.328-343.
De Campos, L.E.P. & Menezes, M.S.S. (1992). A proposed procedure for slope stability analysis in tropical soils. 6th
Int. Symp. On Landslides, Christchurch. Editor D.H. Bell, Balkema, Rotterdam, Vol. 2, pp.1351-1355.
Dearman, W.R. and Fookes, P.G. (1974). Engineering geological mapping for civil engineering practice in the United
Kingdom. Quarterly Journal of Engineering Geology, Vol.7, pp.223-256.
DeBeer, E.E., Goelen, E., Heynen, W.J. and Joustrak. (1988). Cone Penetration Test: International Reference Test
Procedure. Penetration Testing, ISOPT-1. De Ruiter (ed)., Balkema, Rotterdam, pp.27-52.
DeCourt, L., Muromailic T., Nixon, I.K., Schmertmann, J.H. Thorburn, S.T. and Zolkov, E. (1988). Standard
Penetration Test: International Reference Test Procedure. Penetration Testing, ISOPT-1, DeRuiter (ed)., Balkema,
Rotterdam, pp.3-26.
de Groot, M.B. and Stoutjesdijk, T.P. (1997). Undrained stress path of loose sand predicted from dry tests. Canadian
Geotech. J.:34, pp.131-138.
Delage, P., Suraj de Silva, G.P.R., and De Laure, E. (1987). Un nouvel appareil triaxial pour les sols non saturs. 9th
European Conf. on Soil Mech. And Found. Eng., Dublin, Vol.1, pp.26-28.
Delisle, M.C. & Leroueil, S. (2000). Dtection, laide du pizocne, de zones ramollies dans des pentes argileuses et
valuation de leur comportement mcanique. Report GCT-98-23 prepared for the Ministre des Transports du
Qubec, Universit Laval, Qubec.
DeMatos, M. (1986). A sliding consolidation model for rock avalanches. PhD. Thesis, Univ. of Alberta.
Demers, D. (2000). PhD thesis in preparation. Universit Laval, Qubec, Canada.
Demers, D., Leroueil, S. & DAstous, J. (2000). In situ testing in a landslide area at Maskinong, Qubec. Canadian
Geotechnical J. In print.
Dent, J. (1982). A bi-viscous modified Bingham model of snow avalanche motion. PhD thesis, Montana State
University.
Deschamps, R.J. & Leonards, G.A. (1992). A study of slope stability analysis. ASCE Specialty Conf. on Stability and
Performance of Slopes and Embankments II, San Francisco, Vol. 1, pp.267-291
Descoeudres, F. and Zimmerman, T.H. (1987). Three-dimensional dynamic calculation of rockfalls. In Proceedings,
6th International Congress on Rock Mechanics, Montreal, 1:337-342.
Di Maio, C. (1996a). The influence of pore fluid composition on the residual shear strength of some natural clayey
soils. 7th Int. Symp. On Landslides, Trondheim. Vol. 2, pp.1189-1194.
Di Maio, C. (1996b). Exposure of bentonite to salt solution: osmotic and mechanical effects. Gotechnique, Vol.
46(4), pp.695-707.
Dikau, R., Brunsden, D., Schrott, L. and Ibsen, M. (1996). Landslide recognition, Wiley, Chichester.
Dobry, R., I. (1967). Alvare Seismic failure of Chilean tailings dams. J. Soil. Mech. Found. Eng. Div. ASCE, 93, 6,
pp.327-260
Donald, I.B. and Chen Zuyu (1997) Slope Stability Analysis by the Upper Bound Approach: Fundamentals and
Methods. Canadian Geotech. J.. 34:853-862.
Drumright, E.E., Pfingsten, C.W., Lukas, R.G. (1996). Influence of Hammer Type on SPT results. ASCE J.Geotech.
Eng. Vol. 122, No. 7, pp.598-599.
Duncan, J.M. & Dunlop, P. (1969). Slopes in stiff-fissured clays and shales. J. Soil Mech. Found. Engng, ASCE 95,
No. 2, pp.467-492.
Duncan, J.M. (1992). State of the Art Static Stability and Deformation Analysis in Stability and Performance of
slopes and Embankments II, ASCE Geotechnical Special Publication No. 31, pp.222-266.
Duncan, J.M. (1996(a)). State of the Art: Limit Equilibrium and Finite Element Analysis of Slopes. ASCE J. of
Geotech. Eng. Vol. 122, No. 7, pp.577-591.
Duncan, J.M., (1996(b)). Soil slope stability analysis, In Landslides Investigation and Mitigation. Transport.
Research Board, N.R.C. Spec. Report 247, Washington, DC 1996, Chapter 3, pp. 337-371.

Dunnicliff, J. 1995. Keynote Paper: Monitoring and instrumentation of landslides. Proc. Sixth Int. Symp. on
Landslides, Christchurch. Editor DH Bell, Balkema, Rotterdam, pp.1881-1896.
Eckersley, J.D. (1985). Flow slides in stockpiled coal. Engineering Geology, 22, pp.13-22.
Eisbacher, G.H. and Clague, J.J. (1984). Destructive movements in high mountains. Hazard and Management.
Geological Survey of Canada, paper 84-16, 230pp.
Enegren, E.G. and Imrie, A.S. (1996). Ongoing requirements for monitoring and maintaining a large remedial
rockslide. Proc. Seventh Int. Symp. On Landslides, Trondheim, Editor K. Senneset, Balkema, Rotterdam, pp.16771682.
Escario, V. and Juca, J.F.T. (1989). Strength and deformation of partly saturated soils. 12th ICSMFE, Rio de Janeiro,
Vol. 1, pp.43.46.
Escario, V. and Saez, J. (1986). The shear strength of partly saturated soils. Gotechnique, 36(3), pp.453-456.
Eurocode 7. Geotechnical Design (1994). CEN ENV 1997-1
Evans, S.G. and Clague, J.J. (1988). Catastrophic rock avalanches in glacial environments. Proc., 5th International
Symposium on Landslides, Lausanne. Editor C. Bonnard. Balkema, Rotterdam, pp.1153-1158.
Evans, S.G. and Hungr, O. (1993). The assessment of rockfall hazards at the base of talus slopes. Canadian
Geotechnical Journal, 30: 620-636.
Falcetta, J.L. (1985). Un nouveau modle de calcul de trajectories de blocs rocheux. Revue Francaise de
Gotechnique, 30:11-17.
Fell, R. (1994). Stabilisation of soil and rock slopes state of art. North East Asia Symp. and Field Workshop on
Landslides and Debris Flows, Seoul, July 11-16.
Fell, R. (2000). The E.H. Davis Memorial Lecture, Embankment Dams Some Lessons Learnt and New
Developments. Australian Geomechanics, Volume 35, No.1, March 2000, pp.5-48.
Fell, R. and Hartford, D. (1997). Landslide Risk Management, in Landslide Risk Assessment, Editor D. Cruden and
R. Fell, Balkema, Rotterdam, pp.51-110.
Fell, R., Chapman, T.G. and Maguire, P.K. (1991). A model for prediction of piezometric levels in landslides, in
Slope Stability Engineering, Editor R.J. Chandler, Thomas Telford, pp.73-42.
Fell, R., MacGregor, J.P. and Stapledon, D.H.(1992). Geotechnical Engineering of Embankment Dams, Balkema,
Rotterdam. 675pp.
Fenelli, G.B. and Picarelli, L. (1990). The pore pressure field built up in a rapidly eroded soil mass. Canadian
Geotechnical Journal, Vol.27: pp.387-392.
Ferkh, Z. and Fell, R. (1994). Design of embankments on soft clay. XIII Int. Conf. On Soil Mech. And Foundation
Engineering, New Delhi, India, pp.733-738.
Finlay, P.J., Fell, R. and Maguire, P.K. (1997). The relationship between the probability of landslide occurrence and
rainfall. Canadian Geotech J., 34: pp.984-994.
Finlay, P.J., Mostyn, G. and Fell, R. (1999). Landslide risk assessment: Prediction of travel distance. Canadian
Geotech J., Vol. 36, No. 3, 556-562.
Fleming, R.W., Ellen, S.D. and Algus, M.A. (1989). Transformation of dilative and contractive landslide debris into
debris flows - An example from Marin County, California. Engineering Geology, Vol 27, pp.201-223.
Fourie, A.B. (1996). Predicting rainfall-induced slope instability. Proc. Instn. Civ. Engrs. Geotech Eng., Vol.119,
pp.211-218.
Fread, D.L. (1988). The NWS DAMBRK model: theoretical background and user documentation. Unpublished
report by the Hydrologic Research Laboratory, National Weather Service, silver Spring, Maryland.
Fredlund, D.G., Morgenstern, N.R., and Widger, R.A. (1978). The shear strength of unsaturated soils. Canadian
Geotechnical J., 15(3): pp.313-321.
Fredlund, D.G., Vanapalli, S., Xing, A. and Pufahl, D.E. (1995(a)). Predicting the shear strength function for
unsaturated soils using soil-water characteristic curve. 1st Int. Conf. On Unsaturated Soils, Paris, pp.63-69.
Fredlund, D.G., Xing, A., Fredlund, M.D., and Barbour, S.L. (1995(b)). The relationship of the unsaturated soil shear
strength to the soil-water characteristic curve. Canadian Geotechnical J. 32(3), pp.440-448.
Galster, R.W. (1992). The role of engineering geology in slope and embankment stability analysis, in Stability and
performance of slopes and embankments 11. Editors R.B. Seed and R.W. Boulanger. ASCE, Geotechnical
Special Publication, No.31, 70-94.
Geological Society Engineering Group (1972). The preparation of maps and plans in terms of engineering geology.
Quarterly Journal of Engineering Geology, Vol.7, No.3.
Geological Society Engineering Group (1982). Land surface evaluation for engineering purposes. Quarterly Journal
of Engineering Geology, Vol.15, No.4.
Gillon, M.D. and Hancox, G.T. (1992). Cromwell Gorge Landslides a general overview. Proc. Sixth Int. Symp. on
Landslides, Christchurch. Editor D.H. Bell, Balkema, Rotterdam, pp.83-102.
Gillon,M.D., Denton, B.N. and MacFarlane, D.F. (1992). Field investigations of the Cromwell Gorge landslides.
Proc. Sixth Int. Symp. On Landslides, Editor D.H. Bell, Balkema Rotterdam, pp.111-118.

Glastonbury and Fell (2000). Predicting pre and post failure deformations of large landslides. UNICIV Report.
School of Civil and Environmental Engineering, University of New South Wales. (In preparation).
Glastonbury, J. (1999). Preliminary Study of the pre collapse deformation of cut rock slopes. Unpublished report,
School of Civil and Environmental Engineering, University of New South Wales, Sydney.
Glastonbury, J.P. and Douglas, K.J. (2000). Catastrophic Rock Slope Failures Observed Characteristics and
Behaviour. Proceedings, GeoEng2000 Conference, Melbourne, Australia.
Goldsmith,W., (1950). Impact, the theory and practice of colliding solids. Edward Arnold, London, 379 p.
Government of Hong Kong (1972). Final report of the Commission of Inquiry into the Rainstorm Disasters, 1972.
Hong Kong Government Printer, 94pp.
Graham, J.; Crooks, J.H.A. and Bell, A.L. (1983). Time effects on the Stress-strain behaviour of natural soft clays.
Geotechnique 33, No.3, pp.327-340.
Gu, W.H., Morgenstern, N.R. and Robertson, P.K., (1993). Progressive failure in the Lower San Fernando Dam.
ASCE Journal of Geotech. Eng., 119:333-348.
Hanna, T.H. (1985) field instrumentation in geotechnical engineering. Trans. Tech. Pub.
Hanzawa, H. (1989). Evaluation of design parameters for soft clays as related to geological stress history. Soils and
Foundations, 29(2): pp.99-111.
Hardy, R.M., Morgenstern, N.R. and Patton, F.D. (1978). The Garibaldi Advisory Panel Report. B.C. Department of
Highways, Victoria, B.C., Canada (unpublished).
Hawkins, A.B. and Privett, K.D. (1985). Measurement and use of residual shear strength of cohesive soils. Ground
Engineering, November 1985.
Head, K.H. (1980, 81, 85). Manual of soil laboratory testing, Vols. 1 and 3. Pentech Press.
Heim, A. (1932.) Landslides and Human Lives. (Bergsturz and Menchen leben). N. Skermer, Translator. Bi-Tech
Publishers, Vancouver. 196p.
Hendron, A.J. and Patton, F.D. (1985). The Vaiont Slide. U.S. Army Corps of Engineers, Technical Report, Vol.85,
No.5, 104p.
Hendron, A.J. and Patton, F.D. (1985). The Vaiont Slide, a geotechnical analysis based on new geologic observations
of the failure surface. Tech. Report GL-85-5, Department of the Army, US Army Corps of Engineers, Washington
DC (2 volumes).
Hight, D.W., Georgiannou, V.N., Martin, P.L. & Mundegar, A.K. (1998). Flow slides in micaceous snads. Int. Symp.
On Problematic Soils, IS-Tohoku 98, Sendai, Japan. Vol. 2, pp.945-958.
Highter, W.H. and Tobin, R.F. (1980). Flow slides and the undrained brittleness index of some mine tailings, in
Mechanics of Landslides and Slope Stability, edited by S.L. Koh. Engineering Geology. Amsterdam, 16, pp.71-82.
HKIE (1998). Soil nails in loose fill, A preliminary study. Geotechnical Division, Hong Kong Institution of
Engineers.
Holtz, R.D. and Schuster, R.L. (1996). Stabilisation of soil slopes, in Landslides investigation and mitigation,
Transportation Research Board Special Report 247, Editor A.K. Turner and R.L. Schuster, National Academy Press,
Washington DC, pp.439-473.
Holtz, W.G. and Ellis, W., (1968). Triaxial shear characteristics of clayey gravel soils.
Howard, K. (1973). Avalanche mode of motion: implication from lunar examples. Science, Vol.180, pp.1052-1284.
Howe,E., (1909). Landslides in the San Juan Mountains, Colo.: including a consideration of their causes and their
classification. USGS Professional Paper No. 67.
Hs, K.J. (1975). Catastrophic debris streams (sturzstroms) generated by rockfalls. Geological Society of America
Bulletin, Vol. 86, pp.129-140.
Hungr, O. (1981). Dynamics of rock avalanches and other types of slope movements. PhD Thesis, University of
Alberta, 500p.
Hungr, O. (1981(a)). Dynamics of rock avalanches and other types of slope movements. PhD Thesis, University of
Alberta, 500p.
Hungr, O. (1981(b)). Mobility of rock avalanches. Report of the Nat. Research Inst. For Earth Science and Disaster
Prevention, Tsukuba, Japan, No.46, pp.11-20.
Hungr, O. (1990). Mobility of rock avalanches. Report of the Nat. Research Inst. For Earth Science and Disaster
Prevention, Tsukuba, Japan, No.46, pp.11-20.
Hungr, O., (1994). A General Limit Equilibrium Model for 3D slope Stability Analysis, Discussion. Can. Geotech.
Journal., 31:793-795.
Hungr, O. (1995). A model for the runout analysis of rapid flow slides, debris flows and avalanches. Canadian
Geotechnical Journal, 32, pp.610-623.
Hungr, O. (1997). Slope stability analysis. Keynote paper, Procs., 2nd. Panamerican Symposium on Landslides, Rio
de Janiero, 3:123-136.
Hungr, O. (2000). Analysis of debris flow surges using the theory of uniformly progressive flow. Earth Surface
Processes and Landforms, 25:1-13 (in press).

Hungr, O. (2000). Analysis of debris flow surges using the theory of uniformly progressive flow. Earth Surface
Processes and Landforms, 25: pp.1-13. (Not in press any more).
Hungr, O. and Evans, S.G. (1993). The failure behaviour of large rockslides. Geological Survey of Canada Open File
2598pp.
Hungr, O. and Evans, S.G. (2000). Prediction of the failure behaviour of large rockslides. Submitted for publication
to Engineering Geology.
Hungr, O. and Evans, S.G., (1996). Rock avalanche runout prediction using a dynamic model. Procs., 7th.
International Symposium on Landslides, Trondheim, Norway, 1:233-238.
Hungr, O. and Kent, A. (1995). Coal mine waste dump failures in British Columbia, Canada. Landslide News, No.9,
pp.26-27.
Hungr, O., and Evans, S.G., (1997). A dynamic model for landslides with changing mass. In Marinos, P.G., Koukis,
G.C., Tsiambaos, G.C. and Stournaras, G.C., Eds., Procs., IAEG International symposium on engineering geology
and the environment, Athens, June, 1997, 1:719-724.
Hungr, O., Evans, S.G., Bovis, M.J. and Hutchinson, J.N. (2000). A review of the classification of landslides of the
flow type. Submitted to Environmental and Engineering Geoscience.
Hungr, O., Salgado, F.M. and Byrne, P.M., (1989). Evaluation of a three-dimensional method of slope stability
analysis. Canadian Geotech. Journal, 27: 679-686.
Hungr, O., Sun, H.W. and Ho, K.K.S. (1999). Mobility of selected landslides in Hong Kong pilot analysis using a
numerical model. Procs., Hong Kong Institution of Engineers, Geotechnical Division Annual Seminar, May 1999,
pp.169-175.
Hunt, R.E. (1984). Geotechnical Engineering Investigation Manual. McGraw Hill, New York.
Hunter, G. and Khalili, N. (2000). A simple criterion for creep induced failure of over-consolidated clays. Pro.
GeoEng 2000 Conference, Melbourne.
Hunter, G.H. and Fell, R. (2000). The Deformation behaviour of embankment dams. Proc. Eight Int. Symp. On
Landslides, Cardiff, Editor E.H. Bromhead.
Hutchinson, J.N. (1986). A sliding consolidation model for flow slides. Canadian Geotechnical Journal, 23,
pp.115-126.
Hutchinson, J.N. (1987). Mechanisms producing large displacements in landslides on pre-existing shears. Memoir of
the Geological Society of China. No. 9, pp.175-200.
Hutchinson, J.N. (1987). Mechanisms producing large displacements in landslides on pre-existing shears. Memoir of
the Geological Society of China (No.9, December), pp.175-200.
Hutchinson, J.N. (1988). Morphology and geotechnical parameters of landslides in relation to geology and
hydrogeology. Proc. 5th Int. Symp. On Landslides, Lausanne, Editor C. Bonnard, Balkema, Rotterdam, pp.3-35.
Hutchinson, J.N. (1992). Flow slides from natural slopes and waste tips. 3rd National Symposium on Avalanches and
Landslides, La Coruna, Spain, pp.827-841.
Hutchinson, J.N. (1995). Keynote Paper: Landslide Hazard Assessment. Proc. 6th Int. Symp. On Landslides. Editor
D.H. Bell. Balkema, Rotterdam, pp.1805-1841.
Hutchinson, J.N., Brunsden, D. and Lee, E.M. (1991). The Geomorphology of the landslide complex at Ventnor, Isle
of Wight in Slope Stability Engineering, Editor R.J. Chandler, Institution of Civil Engineers, Thomas Telford,
London, pp.213-218.
Hutchinson, J.N., Somerville, S.H. and Petley, D.J. (1973). A landslide in periglacially disturbed Etruria Marl at Bury
Hill, Staffordshire. Quarterly J. Engg. Geol., Vol.6, pp.377-404.
IAEG (1990). IAEG Commission on landslides, suggested nomenclature for landslides. Bulletin of the International
Association of Engineering Geology, No.41, pp.13-16.
Imrie, A.S., Moore, D.P. and Enegren, E.G. (1992). Performance and maintenance of the drainage system at Downie
Slide. Proc. Sixth Int. Symp. On Landslides, Christchurch. Editor D.H. Bell. Balkema, Rotterdam, pp.751-758.
Itasca Consulting Group Inc., (1992). FLAC Version 3.04. Minneapolis, Minnesota.
Iverson, R.M. (1992). Sensitivity of stability analysis to groundwater data. Proc. Sixth Int. Symp. On Landsldies.
Editor D.H. Bell, A.A. Balkema, Rotterdam, pp.451-457.
Iverson, R.M. (1997). The physics of debris flows. Reviews in Geophysics, vol.35, pp.245-296.
Jaeger, J.C., Cook, N.G.W., (1969). Fundamentals of rock mechanics. Methuen.
Jamiolkowski, M.; Ladd, C., Germaine, J.; Lancellotta, R. (1985). New developments in field and laboratory testing.
11th Int. Conf. Soil Mechanics and Foundation Engineering, Vol. 1, Balkema, pp.57-153.
Janbu, N. (1968). Slope Stability Computations. Soil Mechanics and Foundation Engineering Report. Technical
University of Norway, Trondheim.
Jennings, D.N., Newton, C.J., Beetham, R.D. and Smith, G. (1992). Stabilization of the Nine Mile Creek Schist
landslide complex. Proc. Sixth Int. Symp. On Landslides, Christchurch. Editor D.H. Bell, Balkema, Rotterdam,
pp.759-764.
Jeyapalan, K. (1981). Analysis of flow failures of mine tailings impoundments. PhD Thesis, University of
California, Berkeley, 298p.

Johnson, K.A. & Sitar, N. (1990). Hydrologic conditions leading to debris-flow initiation. Canadian Geotechnical
Journal, Vo. 27(6), pp.789-801.
Jorstad, F.A. (1968). Snesmelting som arsak til jordskred pa ostlandet ved manedsskiftet april-mai 1966. Norwegian
Geotechnical Institute, Publication No.75, pp.33-38.
Karlsrud, K., Aas, G. and Gregersen, O. (1984). Can we predict landslide hazards in soft sensitive clays? Summary of
Norwegian practice and experiences. Proc. 4th Int. Sym. Landslides, Toronto. Vol.1, pp.107-130.
Kawamura, K., (1985). Methodology for landslide prediction. Proc. 11th ICSMFE, pp.1155-1158.
Keaton, J.R. and Beckwith, G.H. (1996). Important consideration in slope design, in Landslides Investigation and
Mitigation, Transportation Research Board Special Report 247, Editor A.K. Turner and R.L. Schuster, National
Academy Press, Washington DC, pp.429-438.
Keaton, J.R. and De Graff, J.V. (1996). Surface observation and geologic mapping, in Landslides Investigation and
Mitigation, Special Report 247, Transportation Research Board, Editors A.K. Turner and R.J. Schuster. National
Academy Press, Washington DC, pp.178-230.
Kenny, T.C. & Lau, K.C. (1984). Temporal changes of groundwater pressure in a natural slope of nonfissured clay.
Canadian Geotechnical J. 21(1), pp.138-146.
Khalili, N. and Khabbaz, M.H. (1998). A unique relationship for the determination of the shear strength of saturated
soils. Geotechnique 48, No.5, pp.681-687.
Khalili, N., Fell, R. and Tai, K.S. (1996). A simplified method for estimating failure induced deformation. Proc.
Seventh Int. Symp. On Landslides. Editor K. Sennesett, Balkema, Rotterdam, pp.1263-1268.
Kim, S.K., Hong, W.P. and Kim, Y.M. (1992). Prediction of rainfall-triggered landslides in Korea. Proc. 6th Int.
Symp. on Landslides, Christchurch, Vol. 2, pp.989-999.
King, J., Geotechnical Engineering Office, Hong Kong, (1999). Personal communication.
Kirkebo, S., Nordal, S. & Svano, G. (1996). Time-dependent stability of an excavated slope. 7th Int. Symp. On
Landslides, Trondheim, Vol. 2, pp.1269-1275.
Koerner, H.J. (1976). Reichweite und Geschwindigkeit von Bergsturzen und Fleisschneelawinen. Rock Mechanics,
8, pp.225-256.
Konrad, J.M. (1993). Undrained response of loose compacted sands during monotonic and cyclic compression tests.
Gotechnique, Vol. 43(1), pp.69-89.
Kovacevic, N. (1994). Numerical analyses of rockfill dams, cut slopes and road embankments. PhD. Thesis,
University of London.
Kovari, K. (1990). General report: methods of monitoring landslides. Proc. Fifth Int. Symp. On Landslides.
Lausanne, Vol.3, pp.1421-1434.
Kruhn, M.R. and Mitchell, J.K. (1993). New perspectives on soil creep. ASCE, Journal of Geotechnical
Engineering, Vol 119 (No. 3), pp.507-524.
Kulhawy, F.H. (1992). On the evaluations of static soil properties, in Stability and Performance of Slopes and
Embankments. ASCE Geotechnical Special Publication No. 31, pp.95-115.
Kulhawy, F.H. and Mayne, P.W. (1990). Manual on estimating soil properties for foundation design. Rpt.-6800
Electric Power Research Institute, Palo Alto, 306p.
Lacerda, W.A. (1989). Fatigue of residual soils due to cyclic pore pressure variation. 12th Int. Conf. on Soil
Mechanics and Foundation Engineering, Rio de Janeiro, Vol. 5, pp. 3085-3087.
Ladd, C.C. (1991). Stability evaluation during staged construction (The 22nd Karl Terzaghi Lecture). Journal of
Geotechnical Engineering, ASCE, Vol 117, (No. 4), pp.538-615.
Ladd, C.C. and Foott, R. (1974). New design procedure for stability of soft clays. JASCE Geotech Eng.,GT. 7.
Ladd, C.C., Foott, R., Ishihara, K., Schlosser, F. and Poulos, H.G. (1977). Stress deformation and strength
characteristics. State of art report. 9th Int. Conf. Soil Mech. and Foundation Eng.
Lade, P.V. (1992). State instability and liquefaction of loose fine sandy slopes. ASCE, Geotech J., 118:51-71.
Lade, P.V. (1993). Initiation of static instability in the submarine Nerlerk berm. Canadian Geotechnical J., Vol.
30(6), pp.895-904.
Lade, P.V. and Yamamuro, J.A. (1997). Effects of non plastic fines on static liquefaction of sands. Canadian
Geotechnical J. 34, pp.918-928.
Laflamme, J.F. et al (2000). Simulation of pore pressures of the Saint-Hilaire test excavation. In preparation.
Lafleur, J., Silvestri, V., Asselin, R. and Souli, M. (1988(b)). Behaviour of a test excavation in soft Chaplain Sea
clay. Canadian Geotechnical J., 25(4): pp.705-715.
Lafleur, J., Souli, M. and Silvestri, V. (1988(a)). Pressions interstitielles autour dune fouille exprimentale dans
largile molle. 5th Int. Symp. On Landslides, Lausanne, Vol.1: pp.707-712.
Lam, L. and Fredlund, D.G. (1993). A General Limit Equilibrium Model for Three-dimensional Slope Stability
Analysis. Canadian Geotech. Journ., 30:905-919.
Law, K.T. (1985). Use of field vane tests under earth structures. 11th ICSMFE, San Francisco, Vol.2: pp.893-898.
Law, K.T. and Lumb, P. (1976). A limit equilibrium analysis of progressive failure in the stability of slopes.
Canadian Geotech. J. 15:113-122.

Law, K.T., Shen, J.M. and Lee, C.F. (1998). Strength of loose remoulded granite soil, in Slope Engineering in Hong
Kong. Editors K.S. Li, J.M. Kay, K.K.S. Ho. Balkema, Rotterdam, pp.169-176.
Lebius, J., Robert, J.M. and Rissman, P. (1983). Regional mapping of landslides hazard in Quebec. Symp. On Slopes
in Soft Clays. Linkping, Swedish Geotechnical Institute Report, No.17, pp.205-262.
Lee, H.J., Ellen, S.D. and Kayen, R.E. (1988). Predicting transformation of shallow landslides into high speed debris
flows. Proceedings of the Fifth International Symposium on Landslides, (Bonnard ed.) Lausanne, Switzerland,
Balkema. Vol 1, PP.713-718.
Lee, K.M., Shen, C.K., Leung, D.H.K. and Mitchell, J.K. (1999). Effects of placement method on goetechnical
behavior of hydraulic fill sands. J. Geotech and Geoenvir. Engrg., ASCE, 125(10) 832-846.
Lefebvre, G. (1981). Fourth Canadian Geotechnical Colloquium: Strength and slope stability in Canadian soft clay
deposits. Canadian Geotechnical Journal, Vol 18, pp.420-422.
Lefebvre, G. and LeBoeuf, D. (1986). Rate effects and cyclic loading of sensitive clays. Journal of Geotechnical
Engineering, ASCE, Vol 113, No. 5, pp.476-489.
Lefebvre, G., Pare, J.J. and Dascal, O. (1987). Undrained shear strength in the surficial weathered crust. Canadian
Geotech. J. 24, pp.23-34.
Leroi, E. (1997). Landslide risk mapping: problems, limitations and developments. Procs., Landslide Risk
Workshop, D. Cruden and R. Fell, Eds., Balkema, Rotterdam, pp.239-250.
Leroi, E., Ho, K. and Roberds, W. (2000). Quantitative risk assessment. Proc. GeoEng 2000 Conference, Melbourne
(this volume).
Leroi, E., Landslide risk mapping: problems, limitations and developments. Procs., Landslide Risk Workshop, R.Fell
and D.M. Cruden, Eds., Balkema, Rotterdam , pp. 239-250.
Leroueil, S. & Marques, M.E.S. (1996). State of the Art on the importance of strain rate and temperature effects in
geotechnical engineering. ASCE Convention, Washington, Geotechnical Special Publication No. 61, pp.1-60.
Leroueil, S. (1998) Elements of time-dependent mechanical behaviour of overconsolidated clays. Proc. 51st Canadian
Geotechnical Conf., Edmonton, Vol.2: pp.671-677.
Leroueil, S. (2000). The 39 th Rankine Lecture: Natural Slopes and Cuts, Movement and Failure Mechanisms. To be
published in Geotechnique.
Leroueil, S. and Jamiolkowski, M. (1991) General Report, Session 1: Exploration of soft soil and determination of
design parameters. GeoCoast 91, Yokohama.
Leroueil, S. and Locat, J. (1998) Slope movements: Geotechnical characterisation, risk assessment and mitigation.
Proc. 11th Danube-European Conf. on Soil Mechanics and Geotechnical Engng., Porec, Croatia, pp.95-106. Also
published in Proc. 8th Congress Int. Assoc. Engng. Geology, Vancouver, pp.933-944.
Leroueil, S., La Rochelle, P., Tavenas, F. & Roy, M. (1990). Remarks on the stability of temporary cuts. Canadian
Geotechnical J., Vol. 27, pp.687-692.
Leroueil, S., Locat, J., Vaunat, J., Picarelli, L. and Faure, R. (1996). Geotechnical characterisation of slope
movements. Proceedings of the Seventh International Symposium on Landslides, (Ed. K. Senneset) Trondheim,
Norway, Balkema, Rotterdam. Vol 1, pp.53-74.
Lewis, M.R. and Moore, D.P. (1988). Construction of the Downie Slide and Dutchmans Ridge drainage adits. Proc.
7th Ann. Canadian Tunnelling Conf. Tunnelling Assoc. of Canada. Edmonton Alberta, pp.238-247.
Lied, K. (1977). Rockfall problems in Norway. In Rockfall dynamics and protective works effectiveness, ISMES
Publication No. 90: pp.51-54, Bergamo, Italy.
Lillesand, T.M. (1987). Remote sensing and image interpretation. Wiley, New York.
Li-Tianchi (1983). A mathematical model for predicting the extent of a major rockfall. Zeitschrift fur
Geomorphologie, 27, 4, pp.473-482.
Lumb, P. (1975). Slope failures in Hong Kong. Quarterly Journal of Engineering Geology, 8: pp.31-65.
Lunne, T., Eide, O. and de Ruiter, J. (1976). Correlations between cone resistance and vane shear strength in some
Scandinavian soft to medium stiff clays. Canadian Geotechnical J., 13(4): pp.430-441.
Lunne, T., Robertson, P. and Powell, J.M. (1997). Cone Penetration Testing. Blackie Academic and Professional.
Lupini, J.F., Skinner, A.E. and Vaughan, P.R. (1981). The drained residual strength of cohesive soils. Geotechnique
31, No.2., pp.181-213.
Matouk, A., Leroueil, S., and La Rochelle, P. (1995). Yielding and critical state of a collapsible unsaturated silty
soil. Gotechnique, 34(3), pp.465-477.
MacFarlane, D.F. and Gillon (1996). The performance of landslide stabilization measures. Clyde Power Project,
New Zealand. Proc. Seventh Int. Symp. On Landslides, Trondheim. Editor K. Sennesett. Balkema, Rotterdam,
pp.1747-1758.
MacFarlane, D.F. and Jenks D.G. (1996). Stabilization and Performance of No. 5 Creek Slide, Clyde Power Project,
New Zealand. Proc. Seventh Int. Symp. On Landslides, Trondheim. Editor K. Senneset, Balkema, Rotterdam,
pp.1739-1746.

MacFarlane, D.F., Pattle, A.D. and Salt, G. (1992). Nature and identification of Cromwell Gorge landslides
groundwater systems in Proc. Sixth Int. Symp. On Landslides, Christchurch. Editor D.H. Bell; Balkema,
Rotterdam. pp.509-517.
MacGregor, J.P., Olds, R. and Fell, R. (1990). Landsliding in extremely weathered basalt, Plantes Hill, Victoria.
Proceedings of the Conference on The Engineering Geology of Weak Rock, (Engineering Group of the Geological
Society), Leeds, pp.443-451.
Marchand, G. (1982). Quelques considrations sur le comportement avant rupture des pentes argileuses naturelles.
M.Sc Thesis, Universit Laval, Qubec, Canada.
Martin, D.C. (1993). Time dependent deformation of rock slopes. PhD Thesis, University of London.
McGown, A., Marsland, A., Radwan, A.M. and Gabr, A.W.A. (1980). Recording and interpreting soil macrofabric
data. Geotechnique 30, No.4,
McGuffey, V.C., Modeer, V.A. and Turner, A.K. (1996). Subsurface exploration, in Landslides Investigation and
Mitigation, special report 247, Transportation Research Board. Editors A.K. Turner and R.L. Schuster, National
Academy Press, Washington DC, pp.231-277.
McLellan, P.J. and Kaiser, P.K. (1984). Application of a two-parameter model to rock avalanches in the Mackenzie
Mountains. Procs., 4th Internat. Symposium on Landsldies, Toronto, Vol.1, pp.135-140.
Mesri, G. and Cepeda-Diaz, A.F. (1986). Residual strength of clays and shales. Geotechnique 36, No.2.
Mikkelsen, P.E. (1996). Field instrumentation, in Landslides Investigation and Mitigation, Special Report 247,
Transportation Research Board. Editors A.K. Turner and R.J. Schuster, National Academy Press, Washington DC,
pp.278-316.
Mitchell, J.K. (1993). Fundamentals of Soil Behavior. John Wiley & Sons, Inc.
Mitchell, R.J. and Markell, A.R. (1977). Flow slides in sensitive soils. Canadian Geotech. J. 9: 11-31.
Montgomery, D.R., Dietrich, W.E., Torres, R., Anderson, S.P., Heffner, J.T. & Loagues, K. (1997). Hydrologic
response of a steep unchanneled valley to natural and applied rainfall. Water Resources Research, Vol. 33(1),
pp.91-109.
Moon, A.T. (1992). Stability analysis in stiff fissured clay at Raby Bay, Queensland. Proc. Sixth Int. Symp. On
Landslides. Christchurch, New Zealand. D.H. Bell editor, Balkema, Rotterdam, Vol.2, pp.536-541,.
Moore D.P. and Imrie, A.M. (1992). Stabilization of Dutchmans Ridge. Proc. Sixth Int. Symp. On Landslides,
Christchurch. Editor D.H. Bell. Balkema, Rotterdam, pp.1783-1788.
Morgenstern, N.R. (1978). Mobile soil and rock flows. Geotechnical Engineering, Vol 9, pp.123-141.
Morgenstern, N.R. (1985). Geotechnical aspects of environmental control. Proceedings, ICSMFE, San Francisco, 1:
pp.155-185.
Morgenstern, N.R. (1990). Instability mechanisms in stiff soils and weak rocks. Proc. 10th Southeast Asian Geotech.
Conf. Taipei, pp.27-36.
Morgenstern, N. R. (1992). The Evaluation of Slope Stability-a 25 Year Perspective. In Stability and Performance of
Slopes and Embankments, Geotechnical Special Publication 31, ASCE, New York, 1, pp.1-26.
Morgenstern, N.R. (1995). Keynote paper: The role of analysis in the evaluation of slope stability. 6th Int. Symp. On
Landslides, Christchurch, Editor DH Bell, Balkema, Rotterdam, Vol. 3, pp.1615-1629.
Morgenstern and Ramly (2000. Personal communciation.
Morgenstern, N. R., and V. E. Price. (1965). The Analysis of the Stability of General Slip Surfaces. Geotechnique,
15, (1), pp.79-93.
Morgenstern, N.R., (1985). Geotechnical aspects of environmental control. Proceedings, ICSMFE, San Francisco,
1:155-185
Mostyn, G.R. and Fell, R. (1997). Quantitative and semi-quantitative estimation of the probability of landsliding, in
Landslide Risk Assessment, Editors D. Cruden and R. Fell. Balkema, Rotterdam, pp.297-316.
Mostyn, G.R. and Li (1993). Probabilistic slope analysis state of play. Probabililistic methods in Geotechnical
Engineering. Editors K.S. Li and S.C.R. Lo, Balkema, Rotterdam, pp.89-109.
Mostyn, G.R. and Small, J.C. (1987). Methods of stability analysis, in Soil Slope Stability and stabilization, Editors
B. Walker and R. Fell. Balkema, Rotterdam, pp.71-120.
Muir Wood, D., Jendele, L., Chan, A.H.C. and Cooper, M.R. (1995). Slope failure by pore pressure recharge:
numerical analysis. 11th European Conf. on Soil Mechanics and Foundation Engineering, Copenhagen, Vol.6:
pp.1-8.
Mller, L. and John, K.W., (1963). Recent developments of stability studies of steep rock slopes in Europe. Trans. Soc.
Min. Eng., pp.226.
Mller, L., (1964). The rock slide in the Vaiont valley. Felsmechanik u. Ingenieurgeologie, II. 3-4, pp.148-212.
Murayama, S. and Shibata, T. (1958) On the rheological characters of clay. Disaster Prevention Research Institute,
Kyoto University, Bulletin No. 26, 1-43.
National Research Council (1983). Safety of existing dams, evaluation and improvement. National Academy Press.
Neuzil, C.E. (1993). Low fluid pressure within the Pierre Shale: A transient response to erosion. Water Resources
Research, Vol.29(7): pp.2007-2020.

Nicoletti, P.G. and Sorriso-Valvo, M. (1991). Geomorphic controls of the shape and mobility of rock avalanches.
Geological Society of America Bulletin, Vol.103, pp.1365-1373.
Niewenhuis, J.D. (1991). The lifetime of a landslide. Balkema, Rotterdam.
Nisbet, R.M. (1973). Energy considerations and movements of landslips with special reference to post-rupture
analysis. M.Sc. Thesis, Imperial College, London.
Norem, H., Locat, J. and Schieldrop, B. (1990). An approach to the physics and the modelling of submarine
flowslide. Marine Geotechnology, 9, pp.93-111.
OBrien, J.S., Julien, P.Y. and Fullerton, W.T. (1993). Two-dimensional water flood and mudflow simulation.
Journal of the Hydraulics Division, ASCE, 119, Hy2, pp.244-261.
berg, A.L., and Sllfors, G. (1997). Determination of shear strength parameters of unsaturated silts and sands based
on the water retention curve. Geotechnical Testing J., 20(1), pp.40-48.
Patton, F.D. and Hendron, A.J. (1974). General Report on Mass Movements. Proc. Second Int. Congress of the Int.
Assoc. of Eng. Geology. Sao Paolo. VGR- pp.1-57.
Peck, R.B., (1969). Advantages and limitations of the observational method in applied soil mechanics. Geotechnique
19, 2, pp.171-187.
Perla, R., Cheng, T.T. and McClung, D.M. (1980). A two-parameter model of snow avalanche motion. Journal of
Glaciology, 26, No.94, pp.197-207.
Pfeiffer, T.J. and Bowen, T.D. (1989). Computer simulation of rockfalls. Bulletin of the Association of Engineering
Geologists, 26:135-146.
Picarelli, L. (1991). Resistenza e meccanismi di rottura nei terreni naturali. Convegno dei ricercatori sul tema:
Deformazioni in prossimita della rottura e resistenza dei terreni naturali e delle rocce, Ravello, Vol.2: pp.II.7-II.61.
Piteau, D.R. and Peckover, L. (1978). Engineering of rock slopes. In Landslides, Analysis and Control. National
Academy of Sciences, Nat. Res. Coun., Washington, D.C., Special Rep. 176:192-226.
Poschmann, A.S., Klassen, K.E., Llugman, M.A. and Goodings, D. (1983) Slope stability study of the South Nation
River and portions of the Ottawa River. Ontario Geological Survey Miscellaneous, Paper 112.
Potter, D. (1972). Computational Physics. John Wiley and Sons, London, 400p.
Potts, D.M., Dounias, G.T. and Vaughan, P.R. (1990). Finite element analysis of progressive failure of Carsington
embankment. Geotechnique, Vol 40, (1), pp.79-101.
Potts, D.M., Kovacevic, N. & Vaughan, P.R. (1997). Delayed collapse of cut slopes in stiff clay. Geotechnique,
47(5), pp.953.982.
Pouget, P. and Livet, M. (1994). Relations entre la pluviomtrie, la pizomtrie et les dplacements dun versant
instable (Site de Salldes, Puy-de-Dome). tudes et recherches des Laboratoires des Ponts et Chausses, Srie
Gotechnique GT57, 132p.
Poulos, S.J., Castro, G. and France, J.W. (1985). Liquefaction evaluation procedure. ASCE J. Geotech Eng. Vol.111:
pp.772-792.
Pradel, D. and Raad, G. (1993). Effectof permeability on surficial stability of homogeneous slopes. ASCE J. Geotech
Eng., Vol.119, No.2, pp.315-332.
Puebla, H., Byrne, P.M. and Phillips, R. (1997). Analysis of the CANLEX liquefaction embankment, prototype and
centrifuge models. Can. Geotech, J.,34:641-657.
Reid, M.E. (1997). Slope instability caused by small variations in hydraulic conductivity. ASCE J. Geotech and
GeoEnv. Eng., Vol.123, No.8, pp.717-725.
Rib, H.T. and Laing (1978). Recognition and identification, in Special Report 176: Landslides, Analysis and
Control. TRB, National Research Council, Washington DC, pp.81-111.
Rickenman, D. and Koch, T. (1997). Comparison of debris flow modelling approaches. Proceedings, First
International Conference on Debris-Flow Hazards Mitigation, ASCE, San Francisco, CA, pp.576-585.
Riemer, W. (1995). Keynote paper: Landslides and reservoirs. Proc. Sixth Int. Symp. on Landslides, Christchurch.
Editor D.H. Bell. Balkema, Rotterdam, pp.1973-2004.
Rodine, J.D. and Johnson, A.M. (1976). The ability of debris heavily freighted with coarse clastic materials to flow on
gentle slopes. Sedimentology, 23: pp.213-224.
Romero, S.U. and Molina, R. (1974). Kinematic aspects of Vaiont Slide. Procs., 3rd Congress ISRM, Denver, 2,
pp.865-870.
Royal Swedish Academy of Science (1995). Recommendations for slope stability analysis. Commission Slope
Stability Report 3:95, 4:95, 5:95. Stockholm.
Ryan, T.M. and Call, R.D. (1992). Application of rock mass monitoring for stability assessment of put slope failure.
Rock Mechanics, Proceedings of the 33rd US Symposium, (J.R. Tillerson and W.R. Wawersik ed.) Santa Fe, New
Mexico, Balkema, Rotterdam, pp.221-229.
Saito, M. (1965). Forecasting the time of occurence of a slope failure. Proceedings of the 6th International
Conference on Soil Mechanics and Foundation Engineering, Montreal, Vol. 2, pp.537-541.
Saito, M. and Uezawa, H. (1961). Failure of soil due to creep. Proceedings of the Fifth International Conference on
Soil Mechanics and Foundation Engineering, Paris, Vol. Vol 1, pp.315-318.

Saito, M., (1980a). Evidental study on forecasting occurrence of slope failure. OYO Technical Note TN-38, 1980,
22p.
Saito, M., (1980b). Semi logarithmic presentation for forecasting slope failure. Proc. 3rd Int. Symp. Landslides, Vol.
1, pp.321-324.
Sllfors, G., Larsson, R. and Ottoson, E. (1996). New Swedish national rules for slope stability analysis. Proc.
Seventh Int. Symp. On Landslides, Trondheim, Editor K. Senneset, Balkema, Rotterdam, pp.377-380.
Salt, G. (1988). Landslide mobility and remedial measures. Proceedings of the Fifth International Symposium on
Landslides, (Bonnard ed.) Balkema, Rotterdam, Vol. 1, pp.757-762.
Santos, O.F., Lacerda, W.A. & Ehrlich, M. (1996). Discussion of Collapse of saturated soil due to reduction in
confinement by Anderson & Riemer 1995. ASCE J. of Geotechnical Engineering, Vol. 122(6), pp.505-506.
Sarma, S. K. (1973). Stability Analysis of Embankments and Slopes. Geotechnique, 23, pp.423--433.
Sasitharan, S., Robertson, P.K., Sego, D.C. & Morgenstern, N.R. (1993). Collapse behavior of sand. Canadian
Geotechnical J., Vol. 30(4), pp.569-577.
Sassa, K. (1988). Geotechnical model for the motion of landslides (Special lecture). Procs., 5th Int. Symp. On
Landslides, Lausanne, Editor C. Bonnard, Balkema, Rotterdam, Vol.1, pp.37-56.
Savage, S.B. and Hutter, K. (1986). The motion of a finite mass of granula material down a rough incline. Journ.
Fluid Mechanics, 199, pp.177-215.
Scheidegger, A.E. (1973). On the prediction of the reach and velocity of catastrophic landslides. Rock Mechanics, 5,
pp.231-236.
Schuster, R.L. (1995). Keynote paper: recent advances in slope stabilisation in Proc Sixth Int. Symp. On Landslides,
Christchurch. Editor D.H. Bell. Balkema, Rotterdam, Vol.3, pp.1715-1746.
Schuster, R.L. (1996). Socioeconomic significance of landslides, in Landslides, Investigations and Mitigation,
Transportation Research Board Special Report 247, Editors A.K. Turner and R.L. Schuster, National Academy
Press, Washington DC.
Schuster, R.L. (2000). Dams on Pre-Existing Landslides. Proc. GeoEng 2000 Conference, Melbourne (this volume)
Seed, R.B., Mitchell, J.K. and Seed, H.B. (1990). Kettleman Hills waste landfill slope failure. ASCE Journal of
Geotechnical Engineering, 116:669-690.
Sharpe, C.F.S. (1938). Landslides and related phenomena. Columbia University Press. N.Y..
Silva, A.J. and Brandes, H.G. (1996). Drained creep behavior of marine clays. Measuring and Modelling Time
Dependent Soil Behavior, ASCE, Geotechnical Special Publication No. 61, (S. &. Kaliakin ed.) Washington, D.C.,
pp.228-242.
Silvis, F. and de Groot, MB. (1995). Flow slides in the Netherlands: experience and engineering practice. Canadian
Geotech. J. 32:1086-1092.
Singh, A. and Mitchell, J.K. (1968a). Closure: General stress-strain-time function for soil. Journal of the Soil
Mechanics and Foundations Division, ASCE, Vol 95, (SM6), pp.1526-1527.
Singh, A. and Mitchell, J.K. (1968b). General stress-strain-time function for soil. Journal of the Soil Mechanics and
Foundations Division, ASCE, Vol 94, (SM1), pp.21-46.
Sivathayalan, S. and Vaid, Y.P. (1998). Truly undrained response of granular soils with no membrane penetration
effects. Canadian Geotech J.: 35, pp.730-739.
Skempton, A.W. (1977). Slope stability of cuttings in brown London Clay. Proceedings of the Ninth International
Conference on Soil Mechanics and Foundation Engineering, Tokyo, pp.261-270.
Skempton, A.W. (1985). Residual strength of clays in landslides, folded strata and the laboratory. Geotechnique, Vol
35, (1), pp.3-18.
Skempton, A.W. and Hutchinson, J.N. (1969). Stability of natural slopes and embankment foundations. Procs., 7th
International Conference of Soil Mechanics and Foundation Engineering, Mexica, State of the Art volume, pp.291340.
Sladen, J.A. and Hewitt, K.J. (1989). Influence of placement method on the in-situ density of hydraulic fill sands.
Canadian Geotech J. 26: 453-366.
Sladen, J.A., DHollander, R.D. and Krahn, J. (1985). The liquefaction of sands, a collapse surface approach.
Canadian Geotech. J. 22: 564-578.
Soeters, R. and van Wester, C.J. (1996). Slope instability recognition, analysis and zonation, in Landslides
Investigation and Mitigation, Special Report 247, Transportation Research Board, Editors A.K. Turner and R.L.
Schuster, National Academy Press, Washington DC, pp.129-177.
Soga, K. and Mitchell, J.K. (1996). Rate dependent deformation of structured natural clays. Measuring and
Modelling Time Dependent Soil Behavior, Geotechnical Special Publication No. 61, (S. &. Kaliakin ed.)
Washington, D.C., ASCE, pp.243-257.
Soussa, J. and Voight, B. (1991). Continuum simulation of flow failures. Geotechnique, 41, pp.515-538.
Spencer, E. (1967). A Method of Analysis of the Stability of Embankments Assuming Parallel Inter-slice Forces.
Geotechnique, 17, (1), pp.11-26.

Stapledon, D.H. (1995). Keynote paper: Geological modelling in landslide investigation. Proc. Sixth Int. Symp on
Landslides, Christchurch, 1992. Editor D.H. Bell, Balkema, Rotterdam, pp.1499-1524.
Stark, T.D. and Duncan, J.M. (1991). Mechanisms of strength loss in stiff clays.ASCE. Journal of Geotechnical
Engineering, Vol 117, (1), pp.139-154.
Stark, T.D. and Eid, H.T. (1997). Slope stability analyses in stiff fissured clays.ASCE. Journal of Geotechnical and
Geoenvironmental Engineering, Vol 123, (4), pp.335-343.
Stoutjesdijk, T.P., de Groot, MB. And Lurdenberg, J. (1998). Flow slide prediction method: influence of slope
geometry. Canadian Geotech J. : 35, pp.43-54.
Sun, H.W., Wong, H.N., and Ho, K.K.S. (1998). Analysis of infiltration in unsaturated ground. Proc. Annual
Seminar on Slope Engineering in Hong Kong (Balkema publisher), Hong Kong, pp.101-109.
Tavenas, F. (1984). Landslides in Canadian sensitive clays - A state-of-the-art. Proceedings of the Fourth
International Symposium on Landslides, Toronto, Vol. Vol 1, pp.141-153.
Tavenas, F. and Leroueil, S. (1980). The behaviour of embankments on clay foundations. Canadian Geotechnical J.,
17(2): pp.236-260.
Tavenas, F. and Leroueil, S. (1981). Creep and failure of slopes in clay. Canadian Geotechnical Journal, Vol 18,
pp.106-120.
Tavenas, F., Blanchet, R., Garneau, R. and Leroueil, S. (1978). The stability of stage-constructed embankments on
soft clays. Canadian Geotechnical J., 15(2): pp.283-305.
Tavenas, F., Flan, P., Leroueil, S. and Lebvis, J. (1983) Remolding energy and risk of slide retrogression in sensitive
clays. Proc. Symp. on Slopes on Soft Clays. Undoping, SGI Report No.17: pp.423-454.
Tavenas, F., Leroueil, S., LaRochelle, P. and Roy, M. (1978). Creep behaviour of an undisturbed lightly
overconsolidated clay. Canadian Geotechnical Journal, Vol 15, pp.402-423.
Ter Stepanian, G. (1975) Creep of clay during shear and its rheological model. Geotechnique, Vol 25, (2), 299-320.
Terzaghi, K. (1950): Mechanics of landslides. Berkey Volume, Geological Society of America, New York, 83-124.
Thorne, C.P. (1984) Strength assessment and stability analysis for fissured clay. Geotechnique, Vol.34, No.3, 305-322.
Thurber Engineering (1992). Failure Behaviour of Large Rockslides. Report to The Geol. Surv. Of Canada and BC
Hydro. Thurber Engineering Ltd, Vancouver.
Tian, W.M., Silva, A.J., Veyera, G.E. and Sadd, M.D. (1994) Drained creep of undisturbed cohesive marine sediments.
Canadian Geotechnical Journal, Vol 31, 841-855.
Tika, Th.E. and Hutchinson, J.N. (1999). Ring shear tests on soil from the Vaiont landslide slip surface.
Geotechnique, 49, pp.59-74.
Trak, B. and Lacasse, S. (1996) Soils susceptible to flow slides and associated failure mechanisms. Proceedings of the
Seventh International Symposium on Landslides, (Senneset ed.) Trondheim, Norway, Balkema, Rotterdam, Vol. Vol
1, 497-506.
Troncoso, J.H., Vergara, A., Avendao, A (1993). The scismic failure of Barahona tailings dam. Prco. Third Int.
Conf. Case Histories in Geotech. Eng.
Turner, A.K. and Jayaprakash, G.P. (1996) Introduction, in Landslides Investigation and Mitigation, Special Report
247, Transportation Research Board, Editors A.K. Turner and R.L. Schuster, National Academy Press, Washington
DC, 3-11.
Turner, A.K. and McGuffey, V.C. (1996). Organisation of investigation process, in Landslides Investigation and
Mitigation. Editors A.K. Turner and R.L. Schister. Transportation Research Board Special Report 247. National
Academy Press, pp.121-128.
Ui, T. (1993). Volcanic dry avalanche deposits identification and comparison with non-volcanic debris stream
deposits. Journ. Of Volcanology and Geothermal Research, No.18, pp.135-150.
Uthayakumar, M. and Vaid, Y.P. (1998). Static liquefaction of sands udner multi axial loading. Canadian Geotech J.
: 35, pp.273-283.
Vaid, Y.P. and Eliadorani, A. (1998). Instability and liquefaction of granular soils under undrained and partially
drained states. Canadian Geotech J. : 35, pp.1053-1062.
Vaid, Y.P., Chung, E.K.F. and Kuerbis, R.H. (1990) Stress path and steady state. Canadian Geotech J.. Vol. 27: 1-7.
Vanapalli, S.K., Fredlund, D.G., Pufahl, D.E., and Clifton, A.W. (1996). Model for the prediction of shear strength
with respect to soil suction. Canadian Geotechnical J., 33(3), pp.379-392.
Varnes, D.J. (1982) Time-deformation relations in creep to failure of earth materials. Proceedings of the Seventh
Southeast Asian Geotechnical Conference, Hong Kong, Vol 2, 107-130.
Varnes, D.J., (1978). Slope movement types and processes. In Landslides, Analysis and Control. National Academy of
Sciences, Nat. Res. Cou., Washington, DC., Special Rep. 176: pp.11-33.
Vaughan, P.R. & Hamza, M.M. (1977). Clay embankments and foundations: Monitoring stability by measuring
deformations. Specialty Session 8: Deformation of earth-rockfill dams, 9th Int. Conf. on Soil Mechanics and
Foundation Engineering, Tokyo, pp.37-48.
Vaughan, P.R. (1991) Keynote paper. Mechanics of landslides, in Slope Stability Engineering. Editor R.J. Chandler.
Thomas Telford. 1-12.

Vaughan, P.R. (1994). Assumption, prediction and reality in geotechnical engineering. Geotechnique 44, No.4,
pp.573-609.
Voellmy, A. (1955). Uber die Zerstorungskraft von Lawinen. Schweiz. Bauzeitung, Vol. 73.
Voight, B., (1988). Materials science law applies to time forecasts of slope failures. Proc. Fifth Int. Symp. On
Landslides, Lausanne. Editor C. Bonnard. Balkema, Rotterdam, pp.1471-1472.
Voight, B.A. (1978) Rockslides and Avalanches. Editor B.A. Voight. Elsevier, Amsterdam.
Voight, Y. (1973). Correlation between Atterberg plasticity limits and residual shear strength of natural soils.
Gotechnique, 23(2), pp.265-267.
Vulliet, L & Hutter, K. 1988. Viscous-type sliding laws for landslides. Canadian Geotechnical J. Vol. 25(3), pp.467477.
Vulliet, L. (1986). Modlisation des pentes naturelles en mouvement. Thesis No. 635, cole Polytechnique Fdrale
de Lausanne, Switzerland.
Walker, B.F. and Mohen, F.J. (1987) Groundwater prediction and control, and negative pore water pressure effects, in
Soil Slope Instability and Stabilisation. Editors B.F. Walker and R. Fell, Balkema, Rotterdam, pp.121-182.
Walker, B.F., Blong, R., and MacGregor, J.P. (1987). Landslide classification, geomorphology and site investigations
in Soil Slope Instability and Stabilisation. Editors B.F. Walker and R. Fell, Balkema, Rotterdam, pp.1-52.
Wedage, A.M.P., Morgenstern, N.R. and Chan, D.H., (1998). Simulation of time-dependent movements in Syncrude
tailings dyke foundation. Canadian Geotechnical Journal, 35:284-298
Whitman, R.V. (1984). Evaluating calculated risk in geotechnical engineering. ASCE Journal of Geotechnical
Engineering, 110(2), pp.145-188.
Wieczorek, G.F., Lips, E.W. and Ellen, S.D. (1989). Debris flows and hyperconcentrated floods along the Wasatch
front, Utah, 1983 & 1984. Bulletin of the International Association of Engineering Geologists, 26(2): pp.191-208.
Wong, H.N. and Ho, K.K.S. (1996). Travel distance of landslide debris. Proc. Seventh Int. Symp. On Landslides,
Trondeim, Editor K. Sennessett, Balkema, Rotterdam, pp.417-422.
Woodward Clyde (1998) Lake Eppalock Main Embankment Remedial Works. Unpublished report for Goulburn
Murray Water.
Worsley, P. (1981) Radiocarbon dating: principles, application and sample collection in Geomorphological
Techniques, Editor Goudie et al, George Allen and Unwin.
WP/WL1 (1993) Multilingual Landslide Glossary. Bi Tech Publishers, Richmond, British Columbia, Canada.
Wu, T.H., Tang, W.H. and Einstein, H.H. (1996) Landslide Hazard and Risk Assessment, in Landslides Investigation
and Mitigation, Transportation Research Board Special Report 247, Editors A.K. Turner and R.L. Schuster,
National Academy Press, Washington DC, pp.106-120.
Wylie, .D.C. and Munn, F.J. (1978) The use of movement monitoring to minimise production losses due to put slope
failures. Proceedings of the First International Symposium on Stability in Coal Mining, Vancouver, Miller
Preeman, pp.75-94.
Wylie, D.C. and Munn, F.J. (1978) The use of movement monitoring to minimise production losses due to put slope
failures. Proceedings of the First International Symposium on Stability in Coal Mining, Vancouver, Miller Freeman,
pp.75-94.
Wyllie, D.C. and Norrish, N.I. (1996) Stabilisation of rock slopes, in Landslides Investigation and Mitigation,
Transportation Research Board Special Report 247, Editor A.K. Turner and R.L. Schuster, National Academy Press,
Washington DC, 474-506.
Yamamuro, J.A. and Lade, P.V. (1997). Static liquefaction of very loose sands. Canadian Geotech. J. 34, pp.905917.
Yim, K.P. and Siu, C.K. (1998) Stability investigation and preventive works design for old fill slopes in Hong Kong, in
Slope Engineering in Hong Kong. Editor K.S. Li, J.M. Kay, K.K.S. Ho. Balkema, Rotterdam. 295-304.
Zavodni, Z.M. and Broadbent, C.D. (1980) Slope failure kinematics. CIM Bulletin (April, 1980), pp.69-74

You might also like