Hungr Et Al 2005 - Landslide Travel Distance

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

State of the Art Paper #4

Estimating landslide motion mechanism, travel distance and velocity


O. Hungr
Earth and Oceans Sciences, University of British Columbia, Vancouver
J. Corominas
Department of Geotechnical Engineering and Geosciences, Technical University of Catalonia, Spain
E. Eberhardt
Earth and Ocean Sciences, University of British Columbia, Vancouver

ABSTRACT: An essential part of any landslide hazard or risk assessment is the prediction of the character of
failure and a quantitative estimate of post-failure motion (“runout”) including travel distance and velocity.
The tools available for these tasks are reviewed. At first, phenomenological and analytical methods for pre-
dicting the failure character and timing are shown. The second part of the paper covers empirical methods of
runout prediction and the third part, analytical methods of runout modelling. Although the various method-
ologies are now highly developed, they are not free of potential for error. It is recommended that multiple
methods be used simultaneously and combined with experienced judgment. Each of the analytical techniques
also require careful verification and calibration against field data.

1 INTRODUCTION failure?” Some landslides are slow and ductile, mov-


ing in a continuous or intermittent manner. They
Since the early work of Heim (1932) and Terzaghi may cover long distances (e.g. earth flows), but the
(1950), landslide researchers have striven to better low velocity permits risk reduction action such as
understand and predict catastrophic slope failure. stabilization or evacuation to be taken. Others are
Despite considerable advances in understanding brittle, meaning that after a certain prelude of slow
landslide mechanisms and being able to simulate deformation, or as a result of sudden loading (e.g.
them using numerical models, prediction of the on- during an earthquake), they accelerate and attain ex-
set of extremely rapid motion and the resulting tremely rapid velocities of the order of 5m/s or
propagation (runout) of the slide mass is still ex- faster, exceeding the speed of a running person.
ceedingly difficult. This paper attempts to list and Such landslides are sometimes referred to as “catas-
critically review the main existing techniques and trophic”.
quantitative models. Given the explosion of litera- How do we recognize whether a given potential
ture dealing with this subject in recent years, our landslide can become extremely rapid? The three
treatment is necessarily selective. possible means of answering the question include
The first section summarizes methods of recog- judgmental approach, based on experience and com-
nizing and modelling failure behaviour, with empha- parison with precedents, experimental approach
sis on recognition of phenomena that may lead to ex- based on monitoring, and analytical approach based
tremely rapid failure. The second section deals with on limit equilibrium or stress-strain analysis. Neither
the key empirical techniques for runout prediction, approach is error-proof, and in most cases, special-
and the third with quantitative modelling of land- ists attempt to apply all three.
slide runout. Both landslides originating in soil and
rock are treated. The landslide terminology used de-
rives from Varnes (1978), Cruden & Varnes (1996) 2.2 Judgmental approach, based on landslide
and Hungr et al. (2001). typology
From experience, we know that certain types of
landslides behave in a brittle manner, while others
2 PREDICTION OF FAILURE BEHAVIOUR tend to be ductile. Unfortunately, there is also a large
transitional group that may exhibit either behaviour,
2.1 Prediction methodologies or both in sequence. Nevertheless, a well-designed
The fundamental question connected with landslide typological classification of landslides permits cer-
hazard assessment is “what will be the character of tain distinctions to be made, at least on a preliminary
basis. The following description of typical soil slide kilometer wide floodplain of the Peace River with
trends is based on Hungr et al. (2001). enough speed to raise a violent wave on the opposite
Extra-sensitive (“quick”) clay flow slides are al- shore (Fig. 1). This is an example where clay soften-
ways extremely rapid events, both at initiation, ret- ing produced a material that is very sensitive in its
rogression and during flow-like travel. The same can bulk behaviour.
be said about flow slides in loose saturated sands, Most shallow slides occurring on steep slopes are
often subaqueous. In fact the term flow slide was extremely rapid, simply as a result of cohesion loss
coined by Casagrande (1976) to signify a slide ac- necessary to start failure (Hungr 2003). They usually
companied by liquefaction of a zone of saturated soil involve loose granular veneer overlying stable sub-
at the rupture surface, which invariably leads to strate. Such failures nearly always begin during
catastrophic acceleration. Speculations periodically heavy rain, ensuring perched saturation of the loose
appear in the literature that certain loose, dry fine- layer. As fast movement occurs, soil situated
grained soils can liquefy due to air pressure in the downslope of the initial failure is over-ridden, lique-
pores. However, this process has yet to be conclu- fied by rapid undrained loading and incorporated in
sively demonstrated (e.g. Crosta 2005). Most truly a growing debris avalanche (Sassa 1985). When de-
dry granular flows are slow. Many extremely rapid bris avalanches enter established steep stream chan-
flow slides appear to consist largely of dry or moist nels or gullies, theybecome channelized, incorporate
soil, with liquefaction affecting only a thin saturated further material as well as water and turn into surg-
layer at the base. Such behaviour is observed in ing, extremely rapid debris flows.
well-graded mine waste (Hungr et al. 2002) and
loess. It is possible that the largest landslide disaster
in history, the 1921 series of loess flow slides in the
Loess Plateau, with 280,000 victims, was due
largely to this mechanism (Close & McCormick
1922). Hunter & Fell (2003) summarize conditions
where rapid sliding of soil slopes can be expected.
They show that, based on case studies, soils suscep-
tible to spontaneous liquefaction and extremely
rapid flow sliding span a very wide range of grada-
tions in the silt, sand and gravel classes. To be rapid,
they show that the soils must be contractive.
Sassa (2002) suggested that liquefaction of soil at
the rupture surface may occur as a result of grain
crushing during long-displacement sliding, as the
modified grain size distribution of the crushed soil
allows closer packing, accompanied by pore-
pressure increase. This could explain the spectacu-
lar mobility of many moderately deep-seated flow
slides in residual soil, which probably begin by slid-
ing on relict joints.
Rotational, translational or compound slides in
non - sensitive clay are usually rapid or slower. They
may transform into earth flows and continue moving
at moderate speeds for hundreds of metres. The
highest recorded speed of an earth flow is about 0.1 Figure 1. The Attachie Slide, Peace River, B.C., Canada. The
m/sec (Hutchinson 1974), although speeds in the or- disturbed slope of clay and silt failed suddenly during a rain
der of metres per minute or per hour are more com- storm. The base of the photo is approximately 1800 m wide
mon, even within the course of surging. (Fletcher et al. 2002).
In stiff overconsolidated clays and silts, care must
be taken to account for physical changes caused by
soil disturbance. A striking example of this is the At- Rock slides in stronger rock, usually with a high
tachie Slide on the Peace River, British Columbia, degree of structural control, are extremely rapid due
described by Fletcher et al. (2002). The slope, com- to sudden loss of cohesion. Many become frag-
posed of glacio-lacustrine clays and silts compacted mented and flow-like, forming extremely rapid rock
by an ice sheet and covered by till, had been slowly avalanches (sturzstroms). Of these, rock collapses
failing by compound sliding for many decades. Sud- are especially sudden and rapid, and involve failure
denly, following a period of heavy rain, 7 million m3 of strong rock controlled by a combination of non-
of the disturbed mass liquefied, descended a bedrock systematic joints and intact rock bridges (Hungr &
scarp at the foot of the slope and flowed across the Evans 2004b). In contrast, rotational rock slumps
and some non-rotational compound slides in very cracks or a system of geodetic monuments covering
weak rock are slow, or at best rapid. Block topples, an entire slope, the empirical technique is applied in
in stronger rock and deriving stability from block the same way; surface displacement measurements
equilibrium, are sudden and extremely rapid. Flex- are recorded over time, which are then analyzed for
ural topples, often involving large slopes in weak, accelerations in order to predict catastro-
highly anisotropic foliated rocks, rely on their stabil- phic/impending failure. As such, empirical methods
ity through interlayer friction and tend to be slow. generally overlook the kinematics and causes of
A general assessment of typical failure behaviour failure, relying instead on the surface manifestation
for various types of landslides is given in Table 1. of the instability (e.g. surface displacements).
Of course, such generalizations must be treated with In describing the dynamics of a landslide, Ter-
caution. The best means of assessing a landslide’s zaghi (1950) suggested that many landslides are pre-
potential for catastrophic motion is to compare it ceded by a gradual decrease in the ratio of shearing
with similar case histories whose failure stage has resistance to shearing strength (defined by the factor
already taken place. of safety) which, in turn, involves downward slope
deformations (Fig. 2). Once the factor of safety
reaches unity, slope movements begin to sharply ac-
2.3 Empirical approach based on monitoring celerate and catastrophic failure occurs. For Ter-
Prediction of time to failure can be based on meas- zaghi, the amount of downslope displacement ac-
urements of surface displacements, repeated over commodated through slope creep prior to failure
time. Unlike most numerical analyses, time is ex- (Dfailure in Fig. 2) was related to the thickness of the
plicitly incorporated into the analysis and the ques- basal shear zone, i.e. the zone within which the state
tion of when a given slope might fail is directly ad- of stress approaches the state of failure. If the zone is
dressed. Such empirical approaches however, are very thin, the slope displacements preceding failure
phenomenological (i.e. ‘holistic’) and disregard de- may be on the scale of millimetres. If the potential
tails of the underlying mechanisms while concentrat- sliding surface involves a thicker zone within a ho-
ing on the overall performance of the system. mogeneous soil material, like clay, then the slope
Whether the displacement measurements are made may experience displacements on the scale of metres
using crack extensometers across individual tension before catastrophic failure can occurs.

Table 1. A simple classification of landslides, showing typical ranges of velocities (classification adapted from Varnes 1978, Hungr
et al. 2001, Hungr & Evans 2004b).
TYPE VELOCITY CLASS* COMMENT
ES VS S M R VR ER
SLIDES IN ROCK
Translational (or Wedge) Rock Slide May be slow in very weak rocks
Rotational Rock Slide(Slump) Very weak rock mass
Compound Rock Slide Various types of mechanisms
Rock Collapse Strong rock, joints, rock bridges
FALLS AND TOPPLES
Rock (Debris) Fall Fragmental fall, small scale
Rock Block Topple Single or multiple blocks
Rock Flexural Topple Very weak rock mass
SLIDES IN SOIL
Clay Slump (Rotational) Non- sensitive
Clay Slide (Compound) Non- sensitive
Sand (Gravel, Talus, Debris) Slide Usually shallow
FLOW-LIKE LANDSIDES
Dry Sand (Silt, Gravel, Talus Debris) Flow No cohesion
Sand (Silt, Debris, Peat) Flow Slide Liquefaction involved
Sensitive Clay Flow Slide Quick clay
Debris Avalanche Non-channelized
Debris (Mud) Flow Channelized
Debris Flood High water content
Earth Flow Plastic clay
Rock Avalanche Begins in bedrock
Rock Slide-Debris Avalanche Entrains debris
* Extremely Slow, Very Slow, Slow, Moderate, Rapid, Very Rapid, Extremely Rapid (>5m/sec; Cruden & Varnes 1996).
a modified version of Voight’s semi-empirical time-
dependent failure criterion to provide velocity
thresholds across time spans of 30, 15 and 7 days
(i.e. defining pre-alert, alert and emergency stand-
ings, respectively). The case histories used involved
large deep-seated, creeping-type rockslides, with ve-
locity threshold values differing by an order of mag-
nitude for the different cases (e.g. for the 7-day
emergency alert: 2-12 mm/day for Val Pola, 74-77
mm/day for Vaiont, 207-923 mm/day for Chuqui-
camata). Given the complexity and differences in the
geological and environmental factors contributing to
rock slope failure, Crosta & Agliardi (2003) note
that as a predictor, such empirical methods only give
an order of magnitude prediction of the failure time.
Judgement should be exercised as to the prevailing
external conditions involved in each individual case,
including factors relating to the reliability of the
monitoring network, the complexity of the dis-
placement pattern, and the contributing influence of
precipitation and/or other loading conditions.
Figure 2. Illustration of ground movements that precede a land-
slide, shown as a function of the factor of safety (after Terzaghi
1950).

Indeed, the monitoring of slope movements had


already become standard practice on most mining
and geotechnical projects. In a move towards failure
prediction, focus shifted to the creep behaviour of
materials and the projection of time to failure during
tertiary creep. Saito (1965) proposed that through
the continuous measurement of relative slope dis-
placements, a constant strain rate may be calculated
and compared to the estimated creep rupture life for
that strain rate as determined through laboratory test-
ing. Voight (1989) and Fukuzono (1990) followed
with observations of a linear relation between the
logarithms of the first and second derivatives of
slope displacement (i.e. creep velocity and accelera-
tion). Fukuzono (1990) used this linear relation to
propose a simple means for predicting catastrophic
failure using the inverse mean velocity (Fig. 3). Sev-
eral rheological models have since been forwarded
(linear, exponential, power law), with further differ-
ences between proposed models arising due to the
subjective nature of what constitutes failure (e.g.
Bhandari 1988). Comprehensive reviews of these
methods are provided by Bhandari (1988), Federico
et al. (2002) and Crosta & Agliardi (2003).
The practical application of these methods to fail-
ure prediction requires that warning thresholds be set
with respect to the magnitudes of velocity, or accel-
eration, for a given unstable slope. Salt (1988) pro-
posed a set of empirical alarm levels for large New
Zealand slides in schist, citing downslope velocities
of 50 mm/day (or accelerations of 5 mm/day2).
Figure 3. Method for temporal prediction of slope failure based
Based on the case histories used, the cited alert on inverse mean velocity as calculated from surface displace-
threshold amounted to a forewarning of approxi- ments (after Fukuzono 1990).
mately 10 days. Crosta & Agliardi (2003) proposed
It should be noted that the previously mentioned cal models fail when deviations induced by seasonal
studies all involved back analyses (i.e. hindsight). variations in temperature and rainfall take place.
Reported cases of successful forward prediction are In many displacement-time records, these sea-
few. In one of the better documented cases involving sonal fluctuations, which give rise to changing pore
the Chuquicamata mine in Chile, Kennedy & Nier- pressure conditions, produce a “stick-slip” or epi-
meyer (1970) report the successful extrapolation of sodic creep behaviour (e.g. Bonzanigo et al. 2001).
displacement vs. time plots, based on open pit bench Heim (1932) encountered these difficulties in his
movements (of the F-2 bench), to predict the correct early attempts to interpret accelerating slope move-
date of catastrophic collapse - five weeks in advance ments for failure prediction, incorrectly predicting
of failure. The collapse criterion used in this predic-
catastrophic failure for a rock slope above the town
tion was 6 m of displacement (Fig. 4), a value based
on engineering judgement, displacement data, rock of Kilchenstock in the Swiss Alps – twice! Follow-
mass quality and lessons learned from two previous ing a first incorrect prediction based on slope accel-
minor failures (Voight & Kennedy 1979). In this erations of 5-10 mm/day, which led to the evacua-
sense, the adopted methodology relied heavily on tion of the town (Fig. 5a), Heim (1932) reported that
experience gained over time. Similar successes in “lack of experience” was the reason for being mis-
forward prediction have likewise been documented led. Associating the accelerating slope movements
by Zvelebil (1984) for a toppling failure in sand- with the wetter Fall and their deceleration with the
stone, Azimi et al. (1988) for a rockslide in gypsum, dryer Winter, two years later when velocities in-
Suwa (1991) for a rockslide in tuff, and Hungr & creased to 20-40 mm/day during the dryer July and
Kent (1995) for coal mine waste dump failures. August months, Heim again predicted catastrophic
The complexity of the measured displacement failure followed by a second evacuation of the town,
pattern, and thus the subjectivity and ambiguity in- and again, a second unexpected cessation in slope
volved in its subsequent interpretation, makes defin- movements (Löw 1997).
ing critical thresholds extremely difficult with re- The case of Kilchenstock may have occurred 75
spect to forward prediction. For example, Petley et years ago, but today, the same difficulties in inter-
al. (2002) maintain that since the linear inverse ve- preting and predicting landslide failure from slope
locity relationship used by Fukuzono can be attrib- displacement records arise. In summer 2001, near
uted to stress-induced brittle fracture processes (e.g. Innertkirchen in the Swiss Alps, accelerations ob-
Main et al. 1993), it is only applicable to landslides served through surface displacement measurements
that develop through the generation of a shear plane of a 250,000 m3 unstable rock mass were interpreted
in previously un-failed rock. For failures involving as reaching a critical level (Fig. 5b) thereby requir-
ductile deformation processes or sliding on existing ing the closure of the only highway leading through
planes of weakness, they found that the inverse ve- the mountain pass (the Grimsel Pass). While the
locity follows an asymptotic trend. Crosta & road remained closed and with failure believed to be
Agliardi (2003) found that the linear inverse velocity imminent, a nearby stream with a flow rate of 9000
trend was only applicable to data characterized by litres/minute was diverted down the tension crack
continuous acceleration and invariant to external for a period of 18 days to accelerate and induce fail-
conditions. Their experiences were that such empiri- ure (Gruner 2001).

Figure 4. Extrapolation of displacement vs. time data used to predict a major rock slope failure at the Chuquicamata mine in the
Chilean Andes, and the corresponding record subsequent to the prediction (after Kennedy & Niermeyer 1970).
placement monitoring on which the predictive
analysis is based (1-5 years in many cases) is only a
small snapshot in time with respect to the natural
processes driving the slope to failure. In the case of
an engineered slope, the time over which failure de-
velops closely corresponds to that over which the
measurements are made. In the case of a natural
slope however, creep deformations and slope
movements have been in action for thousands of
years, and have likely encompassed numerous cycles
of acceleration and deceleration. From this, it is dif-
ficult to say whether slope accelerations observed in
a) a short time period are those indicating imminent
failure, or are only a short-term slip interval and sec-
ond-order acceleration (like hundreds before it),
which with time may eventually lead to a first-order
tertiary creep acceleration and catastrophic failure.
The answer to such questions likely lies in the
mechanism of failure, the extent of basal shear plane
development and how much deformation/strain a
given slope can accommodate. As Bhandari (1988)
notes, criteria for failure prediction based on the rate
of slope movement will eventually require that they
be related to the state of a slope prior to failure.

2.4 Numerical approach


b) Mechanistic approaches try to break the problem
down into its constituent parts to understand the
Figure 5. Extrapolation of displacement vs. time data used to cause and effect relationships (and their evolution),
incorrectly predict major rock slope failures in the Swiss Alps: which govern the behaviour of a system. In this
a) at Kilchenstock (after Heim 1932); b) at Innertkirchen with
inset showing velocity vs. time plot and response to different
sense, numerical techniques are commonly em-
attempts to induce failure (after Gruner 2001). ployed to study the balance between driving forces
and resisting forces within a given slope, and the in-
Through this action, slope accelerations increased teraction between the soil/rock mass and external
but failure did not occur (inset Fig. 5b). A decision environmental factors (e.g. pore pressures, seismic
was then taken to use 19 tonnes of explosives to loading, etc.). However, this requires tighter controls
bring down approximately 150,000 m3 of unstable on boundary conditions, material properties and soil-
rock in what would be the largest controlled blast in /rock mass constitutive relationships. Still, through
Switzerland. Following the blast, official figures the use of elasto-plastic yield criteria, numerical
were not released but inspection of before and after methods have been applied to determine/predict the
photos suggested that maybe only 50% of the (sup- location of a potential failure (for example around an
posedly critical) unstable mass was brought down. A open pit mine), the stability state of a slope as a
second blasting campaign was subsequently carried function of changing environmental factors, the
out in August 2002 to bring down the remaining mode of failure, and/or the potential depth/volume
mass (Gruner 2003). of failure. Although time-dependent constitutive re-
Such examples of failed prediction should not be lationships are available (e.g. creep models), nu-
interpreted as a criticism of those involved in the de- merical methods are rarely directed towards tempo-
cision making process. Instead, they demonstrate the ral prediction given the complexity and detailed data
inherent difficulty in relying solely on phenomenol- requirements such predictions would require.
ogical-based analyses in which the underlying kine- Numerical approaches provide a means to ana-
matics, controlling processes and failure mecha- lyze factors relating to landslide behaviour, includ-
nisms are largely ignored. The prevalent use of ing extent, depth and volume of a potential failure,
surface displacement measurements obviously ad- mode of failure and stability state as a function of
dresses certain economic realities in terms of what changing environmental factors. In their simplest
may be feasible for on-site monitoring of a given form, limit equilibrium techniques examine static
slope. Yet it must also be asserted that only so much stability, where a simple balance of disturbing and
can be inferred at surface when the problem itself resisting forces is used, for example, to search for
takes place at depth. Moreover, the time span of dis- the most likely depth of failure (i.e. most critical slip
surface). A comparison of the different solutions to construct the model, and the location of tension
employed to achieve a determinate solution is given cracks at the top of the slope was used to constrain
by Fredlund & Krahn (1977). These methods, how- the model results. The assumption of a continuum,
ever, do not take into account the stress-strain be- although discounting the importance of discontinui-
haviour of the slope mass. Numerical methods on ties in controlling the path of the rupture surface,
the other hand, do utilize stress-strain constitutive provides a clear picture of strain localization and the
relationships to calculate the stresses and deforma- development of a displacement/velocity discontinu-
tions in a slope, as well as the evolution of these de ity delimiting the depth of failure (Fig. 8). As sug-
formations as the slope mass strength degrades and gested by Griffiths & Lane (1999), failure occurs
failure localizes and initiates. “naturally” within the zones of the slope mass where
Numerical methods are generally divided into the shear strength is insufficient to resist the shear
continuum and discontinuum techniques (Fig. 6). stresses, with yielding stresses being redistributed to
Hybrid codes, involving the coupling of these two neighbouring zones. Recent examples involving
techniques, have also been introduced to maximize these techniques applied to case histories of massive
their respective key advantages. Coggan et al. unstable/failed slopes include Benko & Stead
(1998) and Stead et al. (2001) summarize the advan- (1998), Agliardi et al. (2001), Hajiabdolmajid &
tages and limitations inherent in these different Kaiser (2002) and Eberhardt et al. (2004). More ad-
methodologies. The technique(s) chosen depends on vanced numerical methods for modelling shear face
both the site conditions and the potential mode of localization in conjunction with adaptive remeshing
failure, with careful consideration being given to the (i.e. mesh refinement) techniques is discussed by
varying strengths, weaknesses and limitations inher- Zienkiewicz & Huang (1995).
ent in each methodology. Chiriotti et al. (1999) sug- Where the slope instability mechanism is largely
gest that the objectives of a numerical analysis influenced by discontinuities (e.g. joints, faults, bed-
should be directed towards providing a reasonable ding planes, etc.), as is the case for many rock
understanding of the current conditions of the slope, slopes, discontinuum methods provide a more suit-
making it possible afterwards to predict possible able alternative for analyzing landslide behaviour.
evolution scenarios of the instability (with attention
paid to triggering mechanisms which may lead to
catastrophic failure). The degree to which these ob-
jectives can be achieved depends on the quality of
the input data. High quality data enables the objec-
tives to focus more on prediction (i.e. forward mod-
elling of a potential instability), whereas limited data
may restrict the analysis to establishing and under-
standing the dominant mechanisms that may affect
the behaviour of the system (Fig. 7).
Working from a continuum assumption, the ap-
plication of a finite-element (or finite-difference)
analysis does not require any pre-definition of the
failure surface. Instead, a yield criterion (e.g. Mohr-
Coulomb, Hoek-Brown, etc.) can be employed to
Figure 6. Overview of continuum and discontinuum numerical
model and forward-predict the shape and location of methods, with discontinuum model showing cut-away of inter-
the failure surface by following the elastic-plastic acting blocks discretized into deformable elements.
transition of groups of elements as they pass from an
initial linear elastic state to an ultimate state of plas-
tic yield. No commitment is required in the analysis
as to any particular form of the failure mechanism a
priori (Griffiths & Lane 1999). Figure 8 provides an
example for a rock slope in southern Switzerland,
for which a prediction of potential rockslide volume
was required to perform a runout analysis to assess
the risk to sensitive industry infrastructure in the val-
ley below. The model was solved using a strain sof-
tening elasto-plastic constitutive model (decreasing
strength as a function of increasing plastic strain;
e.g. Lo & Lee 1973). The data available for the as- Figure 7. Spectrum of modelling situations and corresponding
sessment were limited to those collected through applicability (after Itasca 2000).
geological mapping and field observations. Through
this, the surface topography and geology were used
Figure 8. Forward prediction of the shape and location of a rock slope failure surface using an elasto-plastic continuum analysis.
Model results show the plasticity indicators for several stages in the progressive development of the failure surface, and below, the
corresponding localization of plastic shear strains.

Discontinuum methods, like the distinct-element


method (Hart 1993), treat the problem domain as an
assemblage of deformable blocks for which complex
non-linear interaction between blocks are solved (i.e.
slip and/or opening/closing along discontinuities).
The method is also capable of modelling the defor-
mation and elasto-plastic yielding of the joint-
bounded intact rock blocks, similar to that discussed
for the continuum techniques. Distinct-element
modelling has been used to investigate a wide vari-
ety of rock slope failure mechanisms including those
ranging from simpler planar/translational mecha-
nisms (Costa et al. 1999, Eberhardt et al. 2005a), to
complex deep-seated sliding/rotation (Chryssan-
thakis & Grimstad 1996, Bhasin & Kaynia 2004),
toppling (Board et al. 1996, Nichol et al. 2002) and
buckling (Stead & Eberhardt 1997). These authors
illustrate the need to consider both intact rock and
joint-controlled displacements in the analysis of
complex rock slope instabilities.
Figure 9 provides an example of a distinct- Figure 9. Distinct-element modelling of complex modes of
element analysis performed for a thinly bedded rock slope failure in thinly bedded weak rock, showing plasticity in-
slope at a western Canadian coal mine, examining dicators and movement of movement vectors of driving and
passive slabs (after Stead & Eberhardt 1997).
the potential modes of failure taken as a function of
the orientation of an undetected cross-cutting dis-
continuity. The major failure mechanisms recog- Further extensions of these methods have worked to
nized include bilinear, ploughing and buckling slab combine both continuum and discontinuum tech-
failures, each involving some form of shear or ten- niques to model intact behaviour, interactions along
sile failure near the toe of the slope followed by pla-
existing discontinuities and the generation of new
nar sliding of the driving slab (Stead & Eberhardt fractures (i.e. the transition from a continuum to a
1997). By comparing the modelled displacement discontinuum). The simulation of softening and
vectors for each failure mode to pit wall slope meas- damage leading to brittle fracturing is accomplished
urements, a prognosis of the mode of failure and using adaptive remeshing techniques, contact search
therefore the best way to remediate the slope may be algorithms and a fracture energy approach controlled
gained.
Figure 10. Hybrid finite-/discrete-element analysis of the 1991 Randa rockslide showing several stages of progressive brittle failure
(after Eberhardt et al. 2002).

by a designated constitutive fracture criterion (e.g. intense rainfall, earthquakes, etc.). When com-
(Munjiza et al. 1995). Such methods provide a bined with precipitation and infiltration records,
means to model the degree of internal fracturing and thresholds may be determined through which predic-
coherency of a failed slope mass, a key considera- tions of depth and relative time of failure can be
tion in any related runout analysis. If the moving made. Collins & Znidarcic (2004) outline a method-
mass fails coherently, higher velocities may accom- ology combining numerical analysis (for seepage in-
pany failure, whereas if the moving mass is ruptured filtration) and limit equilibrium analysis (for stabil-
internally, it will fail block by block, one after the ity state) to quantify failure depth and time in
other, and the travel distance and area covered relation to soil, slope and rainfall parameters. Shou
would be much more limited (Eberhardt et al. & Wang (2003) performed a similar thresh-
2004a). Figure 10 shows a series of model snapshots old/sensitivity analysis in relation to dynamic load-
of the breakdown of a large rockslide mass, subse- ing of a landslide failure triggered by the 1999 Chi
quent to failure initiation, based on Stead & Cog- Chi earthquake in Taiwan.
gan's (2005) “Total Slope Failure” approach. The value of understanding such coupled proc-
Through a better understanding of the mode of esses also extends to decisions regarding mitigation
failure, numerical methods have also been extended measures to be undertaken and prediction of the re-
to the problem of predicting the stability state for a sponse of an unstable slope to such measures.
given slope. In general, this has taken two forms: the Bonzanigo et al. (2001) report the case of a massive
use of strength reduction techniques to determine a 800 million m3 deep-seated creeping landslide in the
factor of safety, and the use of coupled models to southern Swiss Alps, which threatened a local com-
test a slope’s sensitivity to various triggering munity located on the foot of the unstable slide mass
mechanisms. Strength reduction techniques imple- (Campo Vallemaggia). Several different mitigation
ment a procedure whereby the shear strength of the measures were proposed based on competing argu-
soil/rock is reduced until collapse occurs, from ments as to the underlying factors causing the insta-
which a factor of safety is produced by comparing bility. Bonzanigo et al. (2001) used instrumentation
the actual shear strength of the material to the re- records to show that sudden accelerations of the
duced shear strength at failure. This technique has slide mass closely corresponded to pore water pres-
the advantage over limit equilibrium solutions of sure increases exceeding an apparent threshold
automatically finding the critical slip surface as a value. Subsequent modelling results showed that be-
function of the stress state and elasto-plastic yielding tween the different mitigation measures imple-
(Dawson et al. 1999). In addition, the coupled influ- mented, deep drainage through the construction of a
ence of pore pressures and/or dynamic loading on drainage adit provided the only significant benefit,
material yield, and therefore the factor of safety, can resulting in a near immediate reduction in water
be more accurately represented. The modelling of head of 150 m and the almost complete cessation of
nonlinear stress-strain behaviour together with one downslope movement (Fig. 11). The coupled hydro-
or more coupled process also allows for deeper in- mechanical modelling procedure involved solving
sights to be gained into slope instability mechanisms for the steady state pore pressure conditions (repro-
and their sensitivity to different landslide triggers ducing those recorded in boreholes), reduction
Figure 11. Coupled hydro-mechanical distinct-element model of Campo Vallemaggia, showing: a-d) slope velocities prior to and af-
ter simulated opening of the drainage adit (after Eberhardt et al. 2005b).

in shear strength properties along the basal sliding hension of the coupled processes and complex
surface (to initiate movements comparable to those mechanisms driving such instabilities.
recorded in the field), and activation of deep drain-
age through the modelled opening of the drainage
adit. The modelling results showed that in terms of 3 EMPIRICAL METHODS FOR ESTIMATING
fracture permeability, where storativities are low, TRAVEL DISTANCE
large water inflows through drainage are not neces-
sary to achieve significant reductions in head, Several empirical methods for assessing landslide
thereby explaining the relatively small water out- travel distance and velocity have been developed
flows measured from the drainage gallery that led to based on field observations and on the analysis of
scepticism as to the effectiveness of the deep drain- the relationship between parameters characterizing
age solution (Eberhardt et al. 2005b). both the landslide (i.e. the volume of the landslide
In presenting these methods, it should be empha- mass) and the path (i.e. local morphology, presence
sized that elements of field mapping and monitoring, of obstructions), and the distance travelled by the
in situ measurements and laboratory testing must landslide debris. The availability of landslide data
also be included if the overall state-of-the-art is to sets has encouraged the performance of simple sta-
move towards the total assessment or prediction of tistical analyses (bivariate and multivariate analyses)
the rock slope stability state. Currently, an integrated which have produced indexes that are directly or in-
network of displacement, pore pressure and micro- directly related to the landslide mobility. As will be
seismic monitoring devices has been installed at a later shown, empirical indexes are based on simpli-
site in southern Switzerland (the Randa Rockslide fied assumptions and, consequently, they might not
Laboratory), to help in quantifying the spatial and have an evident interpretation. Because of this, the
temporal evolution of such processes and to con- lack of agreement among researchers is not unusual.
strain complex numerical models (Willenberg et al. Methods for predicting landslide runout can be
2002, Eberhardt et al. 2004b). Yet it must always be classified as geomorphologically-based, geometrical
emphasized that numerical modelling is only a tool approaches and volume change methods.
and not a substitute for critical thinking and engi-
neering judgement. Still, the potential exists to use
numerical modelling to build upon empirical meth-
odologies to improve the visualization and compre-
3.1 Geomorphological assessement of landslide mapping these types of deposits, which cover an
runout area of about 550 km2 in the Puget Sound lowland
Field work and photo interpretation are the main close to Seattle, has allowed the delineation of the
sources of the geomorphological analysis for deter- largest lahar originating from Mount Rainier in the
mining the travel distance of landslides. The assess- last 10,000 years, which is known as the Osceola
ment of the extent of both ancient and recent land- Mudflow (Fig. 13). This cohesive lahar, which oc-
slide deposits is the basis for defining future travel curred about 5600 years ago, was at least 10 times
distances. The outer margin of the landslide deposits larger than any other known lahar from Mount Rain-
give an appraisal of the maximum distances that ier (Crandell & Mullineaux 1967, Scott & Vallance,
landslides have been able to reach during the present 1993, Hoblitt et al. 1998) and delineates the scenario
landscape, for a span of time that may last for sev- of the maximum extent that similar events might
eral thousands of years (Fig. 12). reach.
The first constraint that this approach has to over- For these large and ancient landslide events, the
come is the proper identification of the landslide de- limit of the affected zones are well established in the
posits. Thus, in mountain regions, steep slopes of first part of the path but in the far reaches the lack of
formerly glaciated valleys are the source of periodi- continuous outcrops makes the delineation of the
cal rock fall events. Isolated boulders of glacial ori- boundaries more difficult. Because of this, areas lo-
gin (erratic boulders) scattered on the valley floor cated immediately beyond distal hazard zones are
not free of risk because the hazard limits can only be
might be mistakenly attributed to rockfall events.
Some features, like the lithological nature of the approximately located, especially in areas of low re-
boulder, the presence of scratches, rounded edges lief. Uncertainties relating to the source, size, and
and other erosive features, may help to discriminate mobility of future debris flows preclude precise lo-
the gravitational from glacial origin. In alluvial fans, cation of the hazard zone boundaries.
The geomorphological approach does not give
debris flow deposits are often interbedded with tor-
rential (aqueous) laid deposits. Several authors have any clue of the emplacement mechanism. Further-
provided both morphological and textural criteria to more, the slope geometry and the circumstances re-
identify debris flows and their extent (Costa 1984, sponsible for past landslides might have changed.
Jackson et al. 1987). Analysis of outcrops of both Therefore, results obtained in a given place cannot
debris flows and mudflow events have allowed the be easily exported to other localities.
delineation of catastrophic avalanches. For instance,

Figure 12. Boundary of the potential rockfall runout area in Santa Coloma (Principality of Andorra), defined by the line that links
the farthest fallen blocks observed in the field (Copons 2004). Arrows indicate historical rockfall paths and solid circles are large
fallen boulders.
Table 2. Equations for landslide travel distance for Hong Kong
slope failures (Finlay et al. 1999).
Dependent Equation
variable
Cut slope LCI Log L = 0.062 + 0.965 Log H – 0.558 Log (tan δ)
Mean Log L = 0.109 + 1.010 Log H – 0.506 Log (tan δ)
UCI Log L = 0.156 + 1.055 Log H – 0.454 Log (tan δ)
Fill slope LCI Log L = 0.269 + 0.325 Log H + 0.166 Log (V/W)
Mean Log L = 0.453 + 0.547 Log H + 0.305 Log (V/W)
UCI Log L = 0.693 + 0.768 Log H + 0.443 Log (V/W)
Retaining LCI Log L = 0.037 + 0.350 Log H + 0.108 Log (V/W)
wall Mean Log L = 0.178 + 0.587 Log H + 0.309 Log (V/W)
UCI Log L = 0.319 + 0.825 Log H + 0.150 Log (V/W)
Boulder fall LCI Log L = 0.041 + 0.515 Log H – 0.629 Log (tan δ)
Mean Log L = 0.253 + 0.703 Log H – 0.417 Log (tan δ)
UCI Log L = 0.466 + 0.891 Log H – 0.206 Log (tan δ)

Note: H is the vertical drop; δ the slope angle; V the landslide volume,
and W the landslide width. LCI and UCI are the lower and upper 95%
confidence intervals, respectively.

Finlay et al.’s (1999) data was a mixture of good


Figure 13. Extent of two Holocene mudflows from Mount and modest quality information which is reflected in
Rainier, WA established from field reconnaissance of the mud- the large scatter of the predicted travel distances.
flow deposits (modified from Crandell et al. 1979).
Hunter & Fell (2003) revised this work using more
selective good quality data and their recommenda-
tions are to be preferred (Fell, personal communica-
3.2 Geometrical approaches tion). The slope failures showing the largest travel
In this section, travel distance (L) is defined as the distance in the data set are fills, followed by retain-
horizontal projection of the line linking the upper ing walls, cuts and rock falls. The interpretation of
part of the landslide source and the outermost edge this is that many of the fills involved loose, granular
of the landslide deposits (Fig. 14). Finlay et al. materials with contractive behavior during shearing.
(1999), using multiple regression analyses obtained The angle of reach (a) is the angle of the line
several expressions for determining travel distance connecting the highest point of the landslide crown
in cut slopes, fills, retaining walls and boulder falls scarp to the distal margin of the displaced mass (Fig.
(Table 2). These models apply only where debris 14). This angle was defined by Heim (1932), who
runs onto a nearly horizontal surface below. named it the fahrböschung angle. Other names given
by various authors include the reach angle (Coromi-
nas 1996), travel angle (Hungr 1990, Cruden & Var-
nes 1996) and travel distance angle (Hunter & Fell
2003).
What the angle of reach represents has numerous
interpretations. It has been considered as a measure
of the relative mobility of the landslide (Nicoletti &
Sorriso-Valvo 1991, Corominas 1996). Shreve
(1968) called this angle the equivalent coefficient of
friction and Scheidegger (1973) refined this concept,
indicating that for a sliding body, the tangent of the
reach angle is, in fact, the coefficient of friction of
the surface of contact between the sliding mass and
the ground, which is also expressed by the ratio be-
tween the vertical drop H and the horizontal compo-
nent of the runout distance L. Several authors sug-
gest, however, that Scheidegger’s assumption is only
valid in the case of the slope of the line linking the
Figure 14. Geometrical variables: vertical drop (H), travel dis- centres of gravity of both the landslide source and
tance (L), reach angle (α), shadow angle (β), source-talus angle deposit (e.g. Hsü 1975, Voight 1978).
(ψ), substrate angle (γ), and shadow distance (S1) From empirical observations, Heim (1932) ascer-
tained the dependence of the travel distance of a
rock avalanche upon the initial height, the regularity
of the terrain and the volume of the rockslide. A cor-
relation was found between the height of fall and the quently, the key point is to assign an appropriate
distance travelled (Li 1983, Nicoletti & Sorriso- value of angle of reach to the landslide source.
Valvo 1991), but this correlation is difficult to apply However, this is not an evident task.
for practical purposes because the height of fall is Many authors have proposed empirical expres-
not known beforehand except for slopes having a sions based on the inverse relationship between the
flat lying topographical surface below. tangent of the reach angle (H/L) and the landslide
A plot of the tangent of the reach angle (H/L) volume. Initial studies assumed that only large land-
against the landslide volume shows that large land- slides and, particularly, rock avalanches, experi-
slides display lower angles of reach than smaller enced a reduction of H/L with volume increase (i.e.
ones (Scheidegger 1973, Hsü 1975). This is the rea- Scheller 1971, Scheidegger 1973, Li 1983). Further
son why large landslides have been considered as studies with smaller landslides (Hutchinson 1988,
being more mobile. The reach angle of large land- Corominas 1996) and rockfall experiments (Okura et
slides and rock avalanches is smaller than the ex- al. 2000) found a similar correlation with the follow-
pected friction angle of dry broken rock (about 32º). ing form:
This greater mobility has been expressed by the “ex-
cessive travel distance” (Hsü 1975), which is the Log tan α = A + B Log V (3)
length (Le) of the horizontal projection of the runout
distance beyond the point where the line traced from Where A and B are constants and V, the volume. In
the landslide crown dipping at an angle of 32º inter- Table 3 there are several suggested expressions for
sects with the topographical surface. this relationship which correspond to the equation of
H the regression line.
Le = L − (1)
tan 32°
Table 3. Regression equation of Log H/L = A + B Log V, for
Large landslides may have Le values of several different landslide inventories. (*) Volume is expressed in 103
kilometres. Corominas (1996) showed that both m3.
large and small landslides had larger mobility that
Authors A B R
expected using a friction angle of 32º. This angle of Scheidegger 1973 0.624 0.15666 0.82
32º is anyway arguable as many landslide materials Li Tianchi 1983 0.664 -0.1529 0.78
have frictional parameters much less than 32º. How- Nicoletti & Sorriso-Valvo 1991 0.527 0.0847 * 0.37
ever, different mechanisms have been suggested to Corominas 1996 (mean) -0.047 -0.085 0.79
explain this higher mobility. The reader will find de-
tailed discussion on these theories in several re-
search papers (Hungr 1990, Van Gassen & Cruden In some of the landslide sets, the correlation coef-
1989). ficients of the relationship between volume and H/L
The volume dependence of the reach has been are too weak to be used for runout prediction
questioned by several authors for both large land- (Nicoletti & Sorriso-Valvo 1991). Similar attempts
slides (Hsü 1975, Hungr 1990) and small landslides using a population of cuts, fills, retaining walls, and
(Hunter & Fell 2003), and other alternative explana- boulder falls in Hong Kong also found a weak corre-
tions have been proposed (i.e. Davies at al. 1999). lation (R2 < 0.2) and a lot of scatter (Finlay et al.
These works show that there is a lack of agreement 1999).
among researchers, and opposite conclusions have In order to improve the regression equations,
been derived from these simple relations. As a con- Corominas (1996) performed an analysis using more
sequence, the use of the reach angle to determine homogeneous landslide populations. Landslides
travel distance has to be made with care. were split in different groups according to their pre-
When the landslide source and potential landslide dominant mechanism (rock falls and avalanches,
volume are known, the travel distance can be ob- translational slides, debris flows and earthflows) and
tained from the following expression: the characteristics of the path (i.e. unobstructed,
channelled, forested, obstructed by an opposite wall,
H etc). The regression equations show a noticeable im-
L= (2)
tan α provement (Table 4). However, a re-analysis of
these data considering only volumes lesser than 106
In practice, for a given landslide source, the drop of m3, have given poorer correlations (Hunter & Fell
fall (H) is sometimes not a variable known before- 2003). The reasons that explain the scattering are
hand, except for slopes having a flat surface below. manifold: (i) different motion mechanisms; (ii) dif-
In such cases, a graphical solution can be obtained ferent material properties, plus residual strength is
by assuming an angle of reach, for which a line can not considered; (iii) pore fluid pressures are not ac-
be traced from the source; the intersection with the counted for; (iv) simplified morphology and variety
topographic surface will give both H and L. Conse- of path constraints; (v) presence of obstacles, etc.
Table 4. Regression equation of H/L versus landslide volume
for different landslide types and paths (from Corominas 1996). To overcome the constraint of the previous estimate
Landslide Paths A B R2 of both drop of fall and volume, other approaches
type have been proposed (Nicoletti & Sorriso-Valvo
Rockfalls All 0.210 -0.109 0.76 1991) that only require the previous estimation of
Obstructed 0.231 -0.091 0.83
Unobstructed 0.167 -0.119 0.92 the elevation difference (this is very appropriate in
Translational All -0.159 -0.068 0.67 slopes having flat surfaces below). These authors
slides Obstructed -0.133 -0.057 0.76 found that for landslide volumes ranging from 5x106
Unobstructed -0.143 -0.080 0.80 to 1.6x109 m3 the ratio Le/L is usually contained be-
Debris flows All -0.012 -0.105 0.76 tween 0.5 and 0.8. Solving the equation:
Obstructed -0.049 -0.108 0.85
Unobstructed -0.031 -0.102 0.87
Earthflows All -0.214 -0.070 0.65 Le = L –H/tan 32º (4)
Unobstructed -0.220 -0.138 0.91
for this range of values, L falls between 3.2H and
8H.
Due to the large scattering of the plots, the use of The rockfall shadow is the area beyond the toe of
such equations for estimating the expected landslide a talus slope that falling boulders can reach by
travel distance needs to be applied with care because bouncing and rolling. Hungr & Evans (1988) and
the mean values may give optimistic results. Many Evans & Hungr (1993) have used the concept of
landslides will travel far beyond the calculated dis- shadow angle (β) to determine the maximum travel
tance. Instead, it is recommended that the the lower distance of a rockfall. It is defined by the angle of
envelope be used (Fig. 14) and, if enough data is the line linking the talus apex with the farthest block
available, the lines that correspond to the different (Fig. 14). This concept is applied only to fragmental
percentiles (98%, 95%, 90%, etc.) of the spatial rockfalls, which are defined as those events charac-
probability. terized by a more or less independent movement of
In many cases it is relatively easy to model the individual particles generally involving less than 105
uncertainty in the travel distance in the calculation m3, although there is no well defined volume limit.
of the hazard. This can be done by assigning a prob- The application of this method also requires the
ability that the travel distance will be in a certain presence of a talus slope since the shadow angle in
range based on the equations in Tables 4 or 5. The delineated from the talus apex, and the talus toe is
lowest envelope gives the minimum reach angle and used as the reference point beyond which the dis-
that will correspond to the maximum landslide tance travelled by the fallen blocks is determined.
runout (Fig. 15). This seems appropriate for prelimi- The basic concept behind the shadow angle is that
nary studies of runout distance assessment (Ayala et most of the kinetic energy of a rock fragment is lost
al. 2003, Corominas et al. 1990, Corominas et al. in the first impacts on the talus slope and that the fi-
2003). Hunter & Fell (2003) found that the travel nal runout is given by the rolling friction angle,
reach angle correlates reasonably well with the which is obtained by projecting the slope of the en-
downslope angle and the degree of confinement of ergy line from the talus apex. An absolute value of
the path (Table 5). However, their data also exhibits 27.5º has been suggested for the minimum shadow
the wide scatter typical of such correlations. angle (Evans & Hungr, 1993), which is a prelimi-
nary estimation of the maximum rockfall reach.
However, flatter angles sometimes result where the
Table 5. Regression equation of H/L versus tangent of the
downslope angle (α) for landslides from natural slopes with
talus and (or) the substrate is relatively smooth, like
different degree of confinement of the path (Hunter & Fell a glacier or a grass covered surface, with angles ev-
2003). entually reaching 23-24º. Domaas (1994) observed
Paths A B R2 SD even smaller angles, as low as 17º, and found that
Unconfined 0.77 0.087 0.71 0.095 the shadow angle is dependent on the height of the
Partly confined 0.69 0.110 0.52 0.110 talus. The minimum shadow angle is related to the
Confined 0.54 0..27 0.85 0.027 talus height (H2 of the Fig. 14) in the following way:

Domaas (1994), determined the reach angle from β = 0.562H2 + 13.7º (5)
the angle (ψ) of the line linking the rockfall source
with the talus toe (Fig. 14), for three intervals of For talus slopes heights less than 200m, the shadow
height of vertical drop (H): angle obtained was smaller than 25º while for talus
slopes lower than 100m, the shadow angle reduced
H < 200 m α = 0.909 ψ – 8º to as much as16º (Domaas 1994).
200 < H < 300 m α = 0.875 ψ – 3.7º
H > 300 m α = 0.842 ψ – 0.7º
Figure 15. Boundary of the potential rockfall runout area in La Cabrera Sierra (Madrid, Spain), defined by the maximun runout line
(MRL) that links the farthest fallen blocks observed in the field. A safety zone of 100 m width has been traced as well (Ayala et al.
2003).

The shadow angle has been used for zoning of Rockfall susceptible zones were defined based on
hazardous rockfall areas in Andorra la Vella (Co- the spatial distribution of the fallen blocks. Several
pons 2004). The shadow angle (β) has been deter- shadow angle boundaries were traced from the dif-
mined for about one hundred blocks and boulders of ferent talus apexes, taking into account the prob-
recent rockfall events and an inverse relationship has abilities of blocks traveling further. Shadow angles
been found between the tangent of this angle and the of 33º, 32º and 30º correspond to a probability of 1,
volume (V) of the individual fragments. The small- 0.1, and 0.01, respectively, while a lower value of
est shadow angles observed were about 26º. A lower 27.5º was taken from Evans & Hungr (1993) due to
envelope of the minimum shadow angles for the dif- the small number of available cases. These angles
ferent volumes has the following shape: define the boundaries between very high, high, me-
dium, low and very low rockfall susceptibility zones
Log tan β = 0.045 Log V – 0.233 (6) (Fig. 16).

Figure 16. Rockfall susceptible zones , from very high (near the slope) to very low (away from the slope), defined from travel
distance probabilities using the shadow angle (Copons 2004).
The Norwegian Geotechnical Institute (NGI) has 3.5 Performance of the empirical methods
developed a method for estimating the travel dis-
Geomorphological methods are purely empirical,
tance of rockfalls detached from mountain sides without the capability of transferring the results to
(Domaas 1994). This method like the shadow angle other areas of interest. Boundaries are defined based
method can be applied only in places where talus on the presence of previously deposited landslide
slope deposits (scree) have developed. Furthermore, material. In areas of low landslide activity, tracing
the validity of the NGI analysis is restricted to rock- this boundary may be a difficult exercise and subject
fall cases where the average ground surface slope to a high degree of uncertainty and error. Geometri-
beyond the talus toe is less than 12º. Under such cal methods, however, are applicable in many other
conditions, a relationship between the total height of environments. All geometrical methods have in
fall (H) and the distance travelled (S1) by the rock common a large scatter in the empirical relationship
blocks beyond the talus toe (see Fig. 14) was de- between parameters. This fact limits their use as a
rived. The greater the height, the longer S1 will be. predictive tool for landslide travel distance predic-
Again, a lot of scatter appears in such plots. A lower tions except where envelopes defining extreme cases
boundary line including 98% of the cases invento- (maximum extent) are considered or when the num-
ried can be traced to show the worst-case scenario. ber of cases available in the database allows a prob-
The boundary line is expressed by the following abilistic analysis. Despite these constraints, geomet-
equation: rical methods are useful.
The calculation of the reach angle requires the
S1 = aH + b = 0.3065H + 24.1 m (7) measurement of both the starting and end points for
individual landslides. For rockfalls, this constrains
3.3 Prediction of area covered by landslide debris the preparation of a database, as events must be re-
cent enough to allow for the identification of both
Another way to express the spatial susceptibility is
source location and end point, otherwise the deter-
by considering the area affected by the arrival of mination of the angle might be subjected to high de-
landslide debris. A rough proportionality has been gree of uncertainty and error.
found between volume (V) of the landslide debris Prediction of landslide debris travel distances re-
and the area (A) covered by it. Such relationships quires knowledge about future detachment locations
have been observed by Davies (1982) and Li (1983), on the slope. A conservative approach in this cir-
the latter providing the following empirical equa- cumstance may involve taking a minimum reach an-
tion: gle from the top of the slope which would overesti-
mate the runout at a site (Evans & Hungr 1993).
Log A = 1.9 + 0.57 Log V (8) Shadow angle determination does not require the
identification of the rockfall source area, only the lo-
cation of the boulders beyond the talus toe. How-
3.4 Volume-change method for debris flow runout ever, this method can not be applied unless a talus
prediction cone has developed at the base of the rock cliff. For
slopes partly covered by fallen blocks, alternative
The volume change method (Fannin & Wise 2001) procedures must be used.
estimates the potential travel distance of debris flows Using envelopes derived through empirical meth-
by imposing a balance between both the volume of ods are conservative but not unrealistic because they
entrained and deposited mass. The path is subdi- are based on observed cases. In effect, the travel dis-
vided into “reaches”, for which reach length, width tance prediction derived corresponds to the most ex-
and slope are measured. The model considers con- treme observed events that are attained by only a
fined, transitional and unconfined reaches and im- minority of the landslide cases. The conservative na-
poses no deposition for flow in confined reaches and ture of this approach may be very appropriate in pre-
no entrainment for flow in transitional reaches. Us- liminary assessments of landslide susceptibility and
ing the initial volume as input and the geometry of hazard. However, for detailed studies the uncertainty
consecutive reaches, the model establishes an aver- in travel distance should be modelled as described
aged volume-change formula by dividing the vol- above. Ideally, travel distance should also be calcu-
ume of mobilised material by the length of debris lated with numerical analyses calibrated for local
trails. The initial mobilized volume is then progres-
conditions.
sively reduced during downslope flow until the Empirical methods are very simple and travel dis-
movement stops (i.e. the volume of actively flowing tances can be obtained very easily. The main advan-
debris becomes negligible). The results give a prob- tage is their simplicity and that they can be imple-
ability of travel distance exceedance that is com- mented in GIS to delineate the areal extent of
pared with travel distances of two observed events. potential slope failures for susceptibility and hazard
mapping purposes (Ayala et al. 2003, Michael-Leiba
et al. 2003, Copons 2004). However, it should like- models that are too complex contain large numbers
wise be noted that assumptions implicit in these of changeable parameters that cannot be reliably di-
methods are not precise and their statistical scatter is mensioned, permitting almost any desired result to
very large. Also, they do not provide kinematic pa- be obtained. Preferably, calibration analyses should
rameters during the runout process, which are be applied to more than one test case, to ensure that
needed for engineering design. the analysis is repeatable.

4 ANALYTICAL METHODS 4.2 Sliding-block models


In the first known attempt to analyse landslide
4.1 General movement, the physicist E. Müller-Bernet, working
Analytical methods seek to model a moving land- in association with Albert Heim (1932), used the
slide using the physical rules of solid and fluid dy- analogy of a block sliding on a curved path, charac-
namics. Most models are solved numerically, using terised by constant frictional resistance. Figure 17a
some form of finite difference or finite element solu- shows the geometry of a path with local slope angle
tions. The three main groups of solutions include the β and a block with a mass M, moving along the path
block (“lumped mass”) models, two-dimensional surface characterized by a Coulomb friction angle, φ.
models looking at a typical profile of the slide and From the Work-Energy Theorem, the rate of change
of kinetic energy of the block equals the work of the
neglecting the width dimension and three-
dimensional models treating the flow of a landslide net force acting on the block.
over irregular 3-D terrain. Most models (but not all)
belonging to the latter two categories are simplified Mv 2
d( ) = FdL (9)
by integration of the internal stresses in either verti- 2
cal or bed-normal directions to obtain a form of
St.Venant equations. Hydraulicians tend to refer to Here, v is the velocity and the net force, F=Mg(sinβ-
integrated 2-D solutions as 1-D and integrated 3-D cosβtanφ) is the difference between the gravity driv-
as 2-D, which may introduce a degree of confusion. ing force, S and the frictional resistance, T and L is
Herein we will retain the original definitions. the curvilinear distance along the path.
Certain groups of researchers attempt to derive The changing kinetic energy of the block can be
the required constitutive relationships from first graphically expressed by plotting the “energy head”,
principles, using the theory of frictional grain flow v2/2g as a vertical height above the current centre of
(e.g. Savage & Hutter 1989), or mixture flow (e.g. gravity of the block. The locus of the resulting
Iverson 1997). Some attempts have been made re- points was called by Müller-Bernet “energy line”
cently to model landslide motion as the movement and is plotted in Figure 17a.
of discrete particles, without pore fluid (e.g. Camp-
bell 1989). Most models use a semi-empirical ap-
proach called “equivalent fluid method” (Hungr
1995), assigning simple constitutive relationships
judged appropriate for a given material. The validity
and parameters of a given rheological model are de-
termined by back-analysis of case histories similar in
character to the case under consideration. Most
models use a unique, constant constitutive relation-
ship for a given case. Others vary the material prop-
erties for different segments of the path or for se-
lected zones within the moving mass such as the
front and the body of the landslide.
The processes involved in the movement of rapid
landslides are very complex and direct measure-
ments of key variables such as pore-pressure and
viscosity are impossible in full-scale examples.
Thus, any theoretical solution must necessarily con-
tain major simplifications. As is the case with static
stress-deformation analyses in soil and rock mechan-
ics, no analytical solution can be relied on without
being calibrated against field observations. Simple Figure 17. Derivation of the sliding block dynamic equations.
models are preferred, as a limited number of pa- a) Path profile; b) Local slope geometry relationships; c) Force
rameters can more easily be constrained. Many diagram.
From Equation (9): cohesionless granular materials appear always to be
frictional (e.g. Hungr & Morgenstern 1984). There-
Mv 2 fore, the turbulent term is needed only in the pres-
d( ) = Mg (sin β − cos β tan φ ) dL (10) ence of a pore-fluid. In the presence of a pore-fluid,
2
the friction angle φb is mediated by a pore-fluid
From this equation and the geometry of Figure 17b, pressure according to the principle of effective stress
the differential gradient of the energy line, dE, and can thus be much lower than the dry friction an-
equals: gle of the material, φ:

v2 tan φb = (1 − ru ) tan φ (13)


dE = dz − d ( ) = (− tan φ )dx (11)
2g The pore-pressure coefficient, ru is the ratio between
the pore fluid pressure and the total normal stress, as
Thus, as Müller-Bernet had shown, the energy line commonly used in soil mechanics. Provided that ru
of a block moving against constant frictional resis- can be assumed constant for the soil mass under
tance is a line, inclined at an angle equal to the ap- consideration, the right hand side of Equation (13)
plicable friction angle. This result formed the theo- remains of a frictional character, i.e. the shear
retical basis of Heim’s fahrböschung method, as strength component represented by it is proportional
described in Paragraph 3.2 above. The theory also to the total normal stress (Hungr 1995), even though
provides a simple means of estimating the velocity it represents both material friction and pore pressure.
of a frictional slide, from the vertical separation be- Körner (1976) showed that the energy line result-
tween the energy line and the ground. ing from the Voellmy rheology is curved and con-
Of course, representing the mass of a moving cave. Thus, for a given overall displacement of the
landslide by a dimensionless block is a major simpli- block on a path profile, the two-parameter model
fication. In particular, the important effect of lateral predicts lower maximum velocity. This result can be
and longitudinal spreading of the slide mass cannot shown by a re-derivation of the energy line differen-
be accounted for. It should also be pointed out that tial equation. Equation (10) is modified by assuming
the Work-Energy Theorem of Equation (9) should that the “block” is of a constant thickness, h and by
be applied with reference to the centre of gravity of introducing the Voellmy rheology from Equation
the sliding mass, not between the crown of the (12) as:
source area and the toe of the deposit. Physically,
therefore, the block theory and the fahrböschung Mv 2 v2
concept are only crude approximations of the flow d( ) = Mg (sin β − cos β tan φ − Mg )dL (14)
2 ghξ
process for most types of landslides.
Then from Figure 17b:
4.3 Rheological relationships
Körner (1976) demonstrated a useful application of v2 v2
dE = dz − d ( ) = (− tan φ − )dx (15)
the simple block model that helps to clarify the in- 2g ghξ
fluence of various constant and changing rheological
relationships on the dynamic behaviour of land- With two parameters instead of one, the Voellmy so-
slides. He noted that the height of the frictional en- lution is no longer unique for a given landslide dis-
ergy line resulting from typical snow avalanche and placement profile. Many different pairs of the bulk
rock avalanche path profiles often implies unrealisti- friction angle and turbulence coefficient can produce
cally high velocities. He then introduced a two- the same displacement. A unique pair of φb and ξ
parameter frictional-turbulent resistance relationship must be selected so as to match any available data
proposed for snow avalanches by Voellmy (1955). on movement velocity. The largest velocity will be
The resisting stress on the base of a flowing sheet of predicted for the frictional rheology, when the turbu-
fluid material is given by the sum of a Coulomb fric- lence coefficient ξ is very large and the second term
tional and Chézy turbulent term (see also Hungr in Equation (12) can be neglected.
1995): An example back-analysis of a real landslide pro-
file, produced by a numerical solution of Equation
V2 (15), is shown in Figure 18. This is a rock slide –
τ = ρgh cos β tan φb + ρg (12)
ξ debris avalanche with a final volume of over
700,000 m3 that occurred in 1999 in the valley of the
Here, h is the normal thickness of the flow, ρ is den- Nomash River, Vancouver Island, Canada and was
sity and ξ is a turbulence parameter similar to the described by Hungr & Evans (2004a). As shown on
Chézy coefficient, with dimensions of L/T2. Dry the profile in Figure 18a, the landslide descended a
steep 600 m high slope, turned 90˚ in azimuth and
flowed for more than 1.5 km along a gently sloping Figure 19, the initial ru was 1.0 (geostatic value), but
valley bottom. Field observations of superelevation the ratio fell to 0.3 during the 50 seconds of travel.
in bends along the path allowed the true velocity to The energy line is convex upwards, so that the ve-
be estimated at three points along the path, as de- locity required to reach a given displacement is
scribed in Hungr & Evans (2004a). Figure 18b greater than that predicted by the frictional model.
shows three block model solutions. The frictional Thus, the sliding-consolidation models will tend to
solution was run with a φb equal to the fahrböschung over-estimate movement velocities to an even
angle (13.8˚), neglecting the turbulent term. Two greater extent than frictional models. It is also un-
Voellmy solutions are also shown, one with a tan(φb) clear whether unhindered consolidation can exist in
of 0.05 and ξ of 400 m/sec2, the second with 0.04 a rapidly shearing mixture of grains and fluid and, if
and 200 m/sec2. All three solutions produce ap- it does, whether the short time interval occupied by
proximately the same overall displacement of the extremely rapid flow slides allows sufficient time for
block, but they differ in their velocity profiles (Fig. significant consolidation to occur (e.g. Hungr & Ev-
18b). As found by Körner, 1976, the frictional ans 2004a).
rheology overestimates the speed, while the Voellmy Another configuration of the block model was
model provides a much closer fit. used by Sassa (1988) to show the influence of pore-
pressure that changes during the flow as a result of
over-riding and undrained loading of saturated soil
found in the path of the landslide. The model shown
by the dash-dot line in Figure 19 is frictional, based
on the assumption of an initial pore-pressure ratio of
0.0 (dry sliding), changing to a value of 0.87 at the
foot of the steep slope, where saturated valley depos-
its may be over-ridden. The pore-pressure change
appears as a sudden reduction in the slope of the en-
ergy line. The resulting velocity profile falls close to
the values estimated in the field.

Figure 18. Example block model analysis of the Nomash Slide,


using frictional and Voellmy rheologies. a) Path profile and en-
ergy lines. b) velocity profiles. The black dots indicate field es-
timates of velocity by Hungr & Evans (2004a).

Other rheologies have been proposed in the con-


text of the block model. The so-called sliding-
consolidation model (Hutchinson 1986) assumes that
high pore pressure forms during failure in the source
area, as a result of undrained pore-pressure response
or liquefaction. As the slide flows down the path, the
Figure 19. Alternative block model analyses of the Nomash
excess pore pressure acting on the rupture surface Slide, using frictional, sliding-consolidation and ru –controlled
dissipates through consolidation. Eventually, the ef- frictional solutions.
fective stress acting on the sliding surface recovers
sufficiently to cause the slide to stop. This process is
illustrated by the short-dashed line in Figure 19. The Sassa’s undrained loading model can simulate the
solution was obtained using the frictional model limited velocities of real landslides, provided that an
controlled by a φb as for the frictional solution. appropriate sequence of ru’s is chosen. Whether
However, the pore-pressure ratio in Equation (13) rate-dependent (turbulent) flow resistance is also in-
was allowed to dissipate as an inverse exponential volved in the same process is a question that remains
function of the elapsed time of travel, approximating unanswered. Possibly, the development of flow re-
pore-pressure diffusion. In the solution shown in sistance in a slide moving over saturated soil in-
volves both changes of pore-pressure and the ap- 4.5 Integrated 2-D models
pearance of turbulent effects. This can be modelled
Two-dimensional solutions of landslide motion can
by a combination of a frictional model in the steep be derived from the momentum equation for un-
proximal part of the path, followed by a Voellmy steady fluid flow. The basic equation can be derived
segment with a relatively low φb combined with a phenomenologically by considering the dynamic
turbulent term. Such configurations have been tried equilibrium of a column isolated from the flowing
using 2-D and 3-D flow models (e.g. Hungr & Evans sheet, as shown in Figure 20.
2004a, McDougall & Hungr 2005).
One piece of evidence in favour of rate-dependent
resistance is the behaviour of debris flows. Debris
flow surges, although fronted by frictional boulder
accumulations often move for a number kilometres
at moderate speeds, implying the existence of a
normal depth - and rate-dependent rheology – in the
liquefied granular mass behind the front (e.g. Taka-
hashi 1991, Hungr 2000).

4.4 True 2-D models based on the Bingham


rheology
The first analytical models dedicated to the dynam-
ics of landslides were two-dimensional and based on
the Bingham rheology, in which the resisting stress
is determined by the flow depth, h, a constant yield Figure 20. A reference column used in deriving the equation of
motion for the 2D flow.
strength, τy and a linear viscous term with Bingham
viscosity η. An explicit formula for the resisting
stress, τ like Equation (12) cannot be given but the
stress can be related to mean velocity, v, through a The stresses acting on the column are integrated in
third-order equation (e.g. Hungr 1995): the bed-normal direction, assuming that flow lines
are parallel with the base, so that there are no shear
stresses on the sides of the column and the normal
h τ 3
stress increases linearly with depth. In hydraulics,
v= (2τ − 3τ y + y2 ) (16)
6η τ these are known as the shallow water assumptions.
The classic, Eulerian form of the governing equation
Trunk et al. (1986) modified an existing true two- of motion is then:
dimensional finite-difference laminar flow code to
accept two viscosities, one applied up to the level of dv dv τ dh
the yield stress and a smaller one above, to approxi-
χv + = g sin β − − gk cos β (17)
dx dt ρh dx
mate Bingham rheology. The same code, called
BVSMAC, was used by Sousa & Voight (1991) and This is a simple re-statement of Newton’s Second
Voight & Sousa (1994) to model rock avalanches. Law, divided by material density, ρ and flow thick-
The advantage of the true 2-D formulation is the ness, h. The left hand side is the acceleration, con-
ability to derive the vertical velocity distribution and sisting of a convective and local part. v is the mean
the stratigraphy of the deposits. On the other hand, flow velocity. The coefficient Χ is a correction fac-
there is no direct justification for using the Bingham tor, required because the convective part of accelera-
model with materials other than clayey debris (e.g. tion is non-linear with depth. The first term on the
Johnson 1970). Voight & Sousa (1994), as well as right-hand side is the gravity component, G. The
Hungr & Evans (1996) both concluded that the second term is the basal shear resistance, T and the
Bingham model used for rock avalanches tends to third is the pressure gradient, P. In fluid, the constant
overestimate velocities and predict excessive longi- k in the pressure term would equal 1.0, consistent
tudinal spreading of deposits. Recently, however, it with hydrostatic conditions. In a frictional mass, it
was found that a small rock avalanche mobilizing may be a lateral pressure coefficient depending on
and entraining large quantities of plastic clayey col- the internal strain of the flowing mass as discussed
luvium could only be modelled realistically using below.
the Bingham model, in order to duplicate a thin, Equation (17) is derived with reference to bed-
even spreading of the deposits (Geertsema et al. normal thickness. A similar equation could be used
2005). with reference to vertical column boundaries, as is
common in hydraulics (e.g. Iverson 2005). However,
the force diagrams used in solving the equilibrium of
the vertical column become severely distorted on The Savage-Hutter model introduced non-
steep slopes and this may affect the performance of hydrostatic internal tangential stress, based on the
the model. assumption that the flowing mass is frictional, con-
Given the highly unsteady nature of landslide mo- trolled by an internal friction angle φi and undergo-
tion, as well as the need to monitor internal strain ing plastic deformation according to the Rankine
and entrainment of material, it is convenient to re- theory. As a result, the coefficient k in the last term
state Equation (17) in its Lagrangian form, where the of Equation (18) varies typically between approxi-
reference coordinate x is not fixed in space, but trav- mately 0.5 in zones where the flowing mass is
els with the flow. Physically, the Eulerian form stretching on convex path segments and about 4
could be imagined as a set of observations made by when it is being compressed, corresponding to
an observer stationed on the bank of a flowing Rankine’s active and passive conditions respec-
stream. The Lagrangian form results from measure- tively. The ability to model non-hydrostatic internal
ments made by an observer floating on top of the stress is perhaps not essential for fluid-like, fully
flow. Mathematically, the Lagrangian assumption saturated flows such as debris floods or fully-
makes the convective acceleration part disappear: liquefied flow slides. But it is important for the large
group of flow-like landslides that involve relatively
dv τ dh dhv strong frictional masses, moving on a basal liquefied
= g sin β − − gk cos β − (18) or otherwise weakened layer. Thus, it is especially
dt ρh dx h
important for rock avalanches, as without it, the mo-
The last term added on the right hand side of Equa- tion of the slide front cannot be properly predicted
tion (18) is an adjustment for the entrainment of ma- (e.g. Hungr 1995). It is also important for any dry
terial from the path during the flow. Such material granular flows, in nature or the laboratory, that ex-
must be accelerated to the reference frame velocity hibit frictional internal strength.
v, consuming some momentum (Hungr 1995). Hungr (1995) added the possibility to change the
A Lagrangian solution of Equation (18) (without width of the flow according to a user-defined path-
the entrainment term) was developed by Savage & width function. This allows the flow depth to reflect
Hutter (1989) for the frictional rheology. The nu- narrowing or spreading of the flow according to the
merical solution is based on the framework illus- anticipated degree of confinement, producing a
trated in Figure 21. Narrow reference columns, as pseudo-3D solution. This approach neglects momen-
depicted in Figure 21 and numbered i=1 to n, sepa- tum expenditure in the direction perpendicular to the
rate wider “mass blocks”, numbered j=1 to n-1. The flow direction. However, comparisons with 3-D so-
mass blocks remain of constant volume, unless en- lutions and real cases indicate that the effects of this
trainment is specified, thus implicitly satisfying the simplification are not significant, except in the case
continuity equation. The solution is explicit. In each of very abrupt changes of confinement (e.g.
time step, Equation (18) is applied to each reference McDougall & Hungr 2005). Figure 22 shows an ex-
column and the column is displaced forward. When ample of a Voellmy analysis of a rock avalanche
all the columns have assumed new positions, the carried out using the model DAN (“dynamic analy-
mass blocks are re-dimensioned by interpolation to sis”). The case illustrated in this analysis is the right-
obtain new estimates of flow depth. hand streamline of a T-shaped constrained rock ava-
lanche shown in Figure 23. Both the pseudo-3D
analysis and the true integrated 3-D analyses pro-
duced very similar results in terms of runout dis-
tance, debris distribution and velocity.
The frictional Savage-Hutter (1989) model was
verified against controlled small-scale laboratory
experiments involving the flow of sand. Hungr
(1995) added an open rheological kernel, allowing
the resisting stress, τ, to be determined by a number
of alternative rheological formulas, while the inter-
nal shearing of the flowing mass remains frictional.
This allows the user to search for the optimal type of
rheology, as required by the “equivalent fluid” ap-
proach. The resisting stress can thus be determined
Figure 21. Framework for the numerical solution of the St. Ve- from Equation (12) for frictional or Voellmy rheol-
nant Equation in the Lagrangien form (Equation 10). After ogy with pore pressure, from Equation (16) for
Hungr (1995).
Bingham rheology, or using viscous, plastic and
other relationships. The model was verified against
laboratory experiments for the dry frictional, fric-
tional with pore pressure and viscous rheologies.
Verification against the BVSMAC model was made
for the Bingham rheology (Hungr 1995). More re-
cently, the DAN model was calibrated against mul-
tiple case histories of flow slides in mine waste and
a specific type of natural debris avalanches (Hungr
at al. 2002, Revellino et al. 2003). The DAN algo-
rithm has been coded into a spreadsheet form and as
a GIS macro by the Geotechnical Engineering Office
in Hong Kong (H.N. Wong, pers. comm. 2003). A
feature allowing entrainment of path material at a
user-prescribed rate was described by Hungr & Ev-
ans (2004a).
Iverson (1997) extended the Savage-Hutter model
to simulate the behaviour of debris flow surges
Figure 22. An example of a pseudo-3D analysis using the fronted by coarse debris accumulations. In his modi-
model DAN (Hungr 1995). The picture represents an instanta- fication, the bouldery front was simulated by a fric-
neous position of the landslide in an isometric view. This tional mass of a given volume, whose pore-pressure
analysis is of the right-hand streamline of the Eaux Froides ratio was assumed to increase linearly from the lead-
rock avalanche shown in Figure 23. Voellmy rheology, φb=12°,
ξ=200 m/sec2. (Data courtesy Dr. J-D Rouiller, CREALP, ing edge backwards. The front was followed by in-
Switzerland) viscid fluid flow. The model results compared fa-
vourably with experimental debris flows observed in
a large-scale flume, but to date no comparison with
natural debris flow surges has been made. As shown
by Hungr (2000), the flow profile is such compound
models is strongly influenced by the volume of the
frontal accumulation, which is a function of the
granulometry of the debris and the maturity of the
longitudinal sorting of a given surge.

4.6 Three-dimensional fluid dynamics models


Several three-dimensional models have been devel-
oped using the standard Eulerian form of the equa-
tions of motion configured as a 3-D extension of
Equation 17. These solutions are implemented on a
fixed rectangular grid. Besides the local acceleration
a of the fluid, the models also need to keep track of
fluid discharge across the fixed grid boundaries and
the associated momentum fluxes.
“FLOW-2D” (O’Brien et al. 1993) is one of the
few existing models actually used for practical work
(it is a 3-D integrated Eulerian model). It is designed
for analysis of the behaviour of debris flows and de-
bris floods on colluvial fans. A good example of its
use is illustrated by García et al. (2003). The flow
resistance in the model is determined using an em-
pirical rheological relationship similar to the
Voellmy formula (Equation 12), but with the addi-
tion of an additional (“viscous”) term, linear with
velocity. Some attempts to calibrate the rheological
formula through laboratory viscometer testing have
been described in the literature, but in practice, the
b three resistance parameters must be dimensioned
Figure 23. A 3-D analysis of the Eaux Froides rock avalanche through back-analysis. The model does not simulate
in Switzerland, using the method of McDougall & Hungr the motion of the landslide from the source area
(2004a). a) Aerial photo of the landslide, b) Deposit calculated along the path. Instead, the input is in the form of a
by the model. Voellmy rheology, φb=12°, ξ=200 m/sec2. (Data flow hydrograph at the head of a depositional debris
and photo courtesy Dr. J-D Rouiller, CREALP, Switzerland).
fan. From there, the model determines the distribu-
tion of debris over the fan surface, allowing for ob-
structions and pathways such as buildings, roads, parallel strain in the moving mass. This provides re-
channels and bridges. These features make the alistic simulation of the internal stress state, while
model relevant to the determination of flow patterns improving the stability of the solution.
on the surface of a fan, subject to debris flow or de- The second major difference is in the way the
bris flood. No allowance is made for frictional boul- model interpolates flow depth to re-establish the free
der fronts and this may limit the ability of the model surface of the flow at the end of each time step. The
to predict hazard intensity in the channel upstream flow surface interpolation method used by Chen &
of the fan, or near the fan apex. The model is not Lee (2000) was based on the finite element method.
suitable for other types of landslides, such as rock or The finite element mesh is subject to distortion with
debris avalanches, or flow slides. Similar models, large displacements. McDougall & Hungr (2004a)
based on dilatant rheology, have been described by based the interpolation on the smooth particle hy-
Takahashi (1991). drodynamics theory, that permits unlimited distor-
Sassa (1988) developed a 3-D Eulerian model tion as well as separation and joining of flowing
based on his approach of frictional flow resistance, masses. An example analysis of a rock avalanche in
combined with variable pore-pressure coefficient ru. the Swiss Alps is shown in Figure 23.
The pore-pressure coefficient must be mapped over An important feature of up-to-date dynamic mod-
the path area beforehand, as an input function and els is the ability to entrain material from the path of
this largely determines the behaviour of the model. the landslide. Many rock avalanches, although start-
Sassa recommends the use of undrained ring shear ing as dry granular flows, override and erode satu-
tests to determine the ru coefficients applicable to rated surficial soils on the flow path, such as collu-
various substrate material types found along the vial and alluvial or even organic deposits. This
path. The model limitations include the inability to increases their volume, but may also change the
simulate non-hydrostatic lateral pressure, or other basal rheology from that of dry frictional sliding to
types of rheology than frictional. viscous or turbulent slurry flow. In case of debris
flows and debris avalanches, the major part of the
volume in motion may result from entrainment.
4.7 Three-dimensional models with non-hydrostatic Hungr & Evans (2004a) analysed two “rockslide-
lateral stress debris avalanches”, in which entrainment of satu-
All three of the 2-D models reviewed in Paragraph rated debris amounted to as much as 50% of the vol-
3.4 and including non-hydrostatic internal stress, ume. They showed that the process could be simu-
have now been extended into three dimensions. Gray lated successfully using Hungr’s (1995) pseudo-3D
et al. (1999) extended the Savage-Hutter theory of model DAN, together with user-specified quantity of
dry, frictional flow, calibrating it against small-scale entrainment, referred to as “erosion depth”. McDou-
laboratory experiments. Iverson & Denlinger (2001) gall & Hungr (2005) repeated the back-analysis with
also extended the mixture-theory-based model of a 3-D model based on Smooth Particle Hydrody-
Iverson (1997) to 3-D and obtained reasonable simu- namics and obtained similar results, using the same
lation of results from debris flow flume experiments. rheological parameters and entrainment quantities.
As in the 2-D case, the debris flow front is modelled Revellino et al. (2003) calibrated DAN for 19 debris
as a frictional mass with pore-pressure, followed by avalanches in pyroclastic veneer of the Campanian
a viscous flow. Interactions between the fluid phase Region, Italy, using constant entrainment depth of
and solid grains are accounted for using Darcian 1.5 m. More than 80% of the volume of these land-
drag rules, with the implicit assumption that any tur- slides resulted from entrainment. One of the cases
bulent drag can be neglected. The model has not yet was re-analysed with a 3-D entraining model by
been tested against natural debris flow examples. McDougall & Hungr, (2004b). From these studies it
Both Chen & Lee (2000) and McDougall & is clear that the entrainment feature is crucial for
Hungr (2004a) extended the DAN model, using an successful modelling of many landslides.
open rheological kernel in the context of the
“Equivalent Fluid approach.
All of the above models take the Lagrangian ap- 4.8 Simplified models
proach similar to the original Savage-Hutter method, A number of 2-D and 3-D dynamic models are based
resulting from a 3-D extension of Equation (18). The on simplifications of the Equation of Motion, based
reference slices shown schematically on Figure 20 on Kinematic Wave theory that neglects local accel-
become columns, subject to bed-parallel fluid thrust eration (Arattano & Savage 1992) or Cellular Auto-
both in the direction of motion and perpendicular to mata (e.g. Segre & DeAngeli 1995). The latter mod-
it. The main difference between the various models els update the state of a landslide mass in a gridwork
lies in the assumptions used to derive the magnitude of cells over series of time steps. The updating con-
and direction of the internal stresses. The McDou- trols transfer of mass and momentum between adja-
gall & Hungr (2004a) model takes advantage of the cent cells on the basis of pre-determined rules,
Lagrangian framework to keep account of the bed- which are a simplification of the physical concepts
discussed previously in connection with the physical width of the deposit (for 3-D models), travel dura-
models. Since physically-based integrated 3-D dy- tion, spot velocities and spot measurements of the
namic models have now been developed to a high flow depth or of the thickness of deposits at certain
degree and computing power is easy to obtain, there locations. Each calibration study should use more
would seem to be less reason at present to use or than one landslide of a given type, in order to prop-
continue developing simplified models. They are erly constrain the model.
beneficial for analysis of sediment transport, where Spot velocities are often estimated using hydrau-
the constitutive relationships are very complex (e.g. lic formulas for superelevation in bends, or runup
Pitman et al. 2003). against adverse segments of the path (e.g. Hungr &
Evans 2004a). At other times, eyewitness estimates
or film/video recordings can be used. Velocities
4.9 Verification and calibration of models have also been back-calculated from observed dam-
Verification testing serves to demonstrate that a age to structures along the path (e.g. Faella & Nigro
given model is physically correct, consistent with (2001). During such back-calculation, care must be
the assumptions of a given method. The best way to taken to correctly specify the mechanism causing the
achieve this is to compare the model results to con- structural damage. For example, much lower veloci-
trolled laboratory experiments utilizing materials ties are needed to damage structures through boulder
with known constitutive relationships. Real landslide impact, then through fluid thrust alone (Revellino et
cases are not as useful for model verification since al. 2003).
neither the constitutive laws nor the details of mo-
tion are sufficiently well known for them.
Many researchers conduct specially designed ex-
periments for this purpose, or adopt detailed experi-
mental results reported by others (e.g. Hutter & Sav-
age 1989, Hungr 1995, Iverson 1997, Denlinger &
Iverson 2001, Iverson et al. 2004, Gray et al. 1999).
For example, Figure 24 shows the analysis by Hungr
(1995) of a dam-break experiment involving the vis-
cous laminar flow of oil. The laboratory experiment
was carried out and reported on by Jeyapalan (1981).
Figure 25 shows the results of a simple verification
experiment on “deflected flow” of dry sand, carried
out by McDougall & Hungr (2004a). These experi-
ments show the model’s ability to steer the flowing
mass along a curving path and also to simulate lat-
eral spreading, controlled by strain-dependent inter-
nal stresses. Iverson et al. (2004) used laser devices
to produce contour maps of moving sand masses, to
better compare with calculated dimensions.
It is important that each analytical model should
be similarly compared with controlled experiments
to confirm the validity of the numerical code, before
attempting calibration against real landslide cases or
use for prediction.
Any dynamic analysis of a landslide necessarily
incorporates important simplifications in terms of
material and pore-pressure spatial heterogeneity as
well as variability of key parameters with time. As is
the case with static analysis of stress in soil or rock,
to be practically useful, models must be calibrated
by comparison with full-scale behaviour of real
events. Provided that a model is correct and veri-
fied, calibration is needed to 1) select the appropriate
rheological kernel and 2) dimension the controlling
resistance parameters. The goal of calibration is to
compare certain control parameters calculated by the Figure 24. A flume test with the laminar flow of oil by Jeya-
model with their equivalents available from observa- palan (1981), back-analysed by Hungr (1995).
tions. The control parameters may include total dis-
placement (“runout”) of the landslide toe, length and
Figure 25. Comparison of actual and predicted flow in a deflected flow experiment using polystyrene beads (McDougall & Hungr
2004a).

In order to make the maximum use of available Only a few systematic calibration studies, using
observations, the calibration needs to be carried out multiple events have been published. Hungr et al.
opportunistically using model results wherever field (2002) found that large flow slides in coal mine
numbers are available. As an example, Hungr & Ev- waste usually conform to the frictional model (Equa-
ans (1996) reported back-analyses of 23 well-known tion 13), with a bulk friction coefficient averaging
large rock avalanches, using Hungr’s model DAN 21°, with a standard deviation of about 3˚. An excep-
(1995) with three alternative rheological kernels: tion occurs in locations where the slide is able to en-
frictional, Voellmy and Bingham. The model pa- train saturated material from the path, in which case
rameters were adjusted in each case so as to obtain a Voellmy analysis is needed and much longer travel
perfect fit in terms total runout distance and the best distances occur. Hungr & Evans (1996) found that
possible fit in terms of velocities. As shown in Fig- 70% of cases of large rock avalanches conform rea-
ure 26, a good fit could only be obtained with the sonably well to the Voellmy model, with a tan (φb)
Voellmy rheology. Both frictional and Bingham of 0.1 and a ξ of 500 m/sec2. Of the remaining 30%
consistently overestimated the velocities. The same some were relatively dry rock avalanches, behaving
study found that the Voellmy rheology produces the as frictional fluids. The remainder were more mobile
best approximation of deposit length and thickness; due to major entrainment of snow, ice or loose satu-
it tends to concentrate the deposits forward into the rated soil and were analysed by the Voellmy model
distal zone. In contrast, the frictional rheology pre- with lower bulk friction angles.
dicts thin tapering fronts (Fig. 27). This is appropri-
ate only for dry, granular landslides and model tests.

100

80
CALCULATED (m/s)

Voellmy
60
Frictional
Bingham
40

20

0 20 40 60 80 100
FIELD VELOCITY (m/s)
Figure 26. Comparison between calculated and observed flow velocities at various locations on the paths of 23 rock avalanches
back-analysed by Hungr & Evans (1996).
2000
calibration runs must be judged on the basis of a suf-
ELEVATION (m) 1800
ficient number of control quantities(e.g. runout dis-
1600 tance, velocities, depths) to securely constrain the
1400 changeable input parameters as well as the overall
1200 design of the model. Wherever possible, calibration
1000 should use more than one event. An example of a
1000 2000 3000 4000
successful calibration was presented by Revellino et
2000
DISTANCE (m)
al. (2003), who showed that successful back-analysis
1800
of 19 similar debris avalanches could be made with a
ELEVATION (m)

single set of Voellmy parameters. A model thus


1600
calibrated is ready for forward predictions.
1400

1200

1000 5 CONCLUSION
1000 2000 3000 4000
DISTANCE (m)
Thanks to intensive research by many groups over
Figure 27. longitudinal debris distribution of the Frank Slide, the last 20 years, numerous powerful tools now exist
Alberta, Canada, calculated by a) frictional and b) Voellmy
rheology ( Hungr 2005). for prediction of the failure behaviour of extremely
rapid landslides. What is most urgently needed,
however, is to establish a connection between field
Revellino et al. (2001) found that 19 of the large observations and analysis. We must try to bridge the
debris flows and debris avalanches derived from py- frequent gap between model developers, many of
roclastic soils in the Campania region of Italy be- whom are trained mainly in theoretical sciences, and
haved as Voellmy fluids, with a tan (φb) of 0.07 and field geologists and engineers who understand earth
a ξ of 200 m/sec2. Ayotte & Hungr (2001) found that structures, materials and processes. Development of
many debris avalanches in Hong Kong, involving focused typological classification of landslide and
both natural and filled slopes, could be simulated systematic, quantitative back-analysis using various
with the frictional model, with a bulk friction angle models is important. Rules must be established for
in the range of 20˚ to 30˚. For those events smaller using sophisticated models and judging their reli-
than 1000 m3, the angle increased to as much as 43˚ ability. Different types of models should be com-
and was inversely related to volume. Debris flows in pared against each other. Above all, we must exer-
the same region conformed better to the Voellmy cise healthy skepticism regarding the results of any
model. The Bingham model appears suitable only model, no matter how complex and impressive.
for landslides containing plastic material (Johnson
1970, Geertsema et al. 2005).
However, much more work is required before we ACKNOWLEDGMENT
can generalize the optimal selection of rheological
parameters for prediction. For the time being, it is The authors wish to acknowledge their colleagues
necessary to conduct focused calibration for each who helped contribute to the results presented here-
given case, by systematically back-analysing exam- in. Prof. Doug Stead (Simon Fraser University) and
ples similar in scale and geological character, con- Dr. John Coggan (University of Exeter) contributed
centrating on the nature of the substrate in the flow to the hybrid FEM/DEM analysis. Scott McDougall,
path. Recently, attempts have been made to calibrate an NSERC Canada scholarship doctoral student at
entraining models, where the rheology of the land- the University of British Columbia and Nikolai
slide changes along the path (e.g. Hungr & Evans Hungr have helped to produce many of the runout
2004a). Such models are more difficult to constrain, modelling results shown here.
as the point of the beginning of entrainment, where
the rheology changes, and the entrainment depth,
become additional input parameters. Fortunately, REFERENCES
surficial geological observations can be used to es-
Agliardi, F., Crosta, G. & Zanchi, A. 2001. Structural con-
timate these parameters, but much more work is re- straints on deep-seated slope deformation kinematics. Engi-
quired in this aspect. neering Geology 59(1-2): 83-102.
In general, reological relationships defining the Aratano, M. & Savage, W.Z. 1992. Kinematic wave theory for
equivalent fluid should be simple with few change- debris flows. USGS Open File Report 92-290, Denver,
able parameters that can be easily constrained. The Colorado.
recommended maximum number of parameters in Ayala, F.J., Cubillo, S., Álvarez, A., Domínguez, M.J., Laín,
L., Laín, R. & Ortíz, G. 2003. Large scale rockfall reach
the basal resistance equation is two. The success of susceptibility maps in La Cabrera Sierra (Madrid) per-
formed with GIS and dynamic analysis at 1:5,000. Natural Costa, J.E. 1984. Physical geomorphology of debris flows. In
Hazards, 30: 325-340. Costa & Fleisher (eds.). Developments and Applications of
Azimi, C., Biarez, J., Desvarreux, P. & Keime, F. 1988. Fore- Geomorphology. Springer Verlag. pp. 268-317
casting time of failure for a rockslide in gypsum. In Bon- Costa, M., Coggan, J.S. & Eyre, J.M. 1999. Numerical model-
nard (ed.), Proceedings of the Fifth International Sympo- ling of slope behaviour at Delabole slate quarry. Interna-
sium on Landslides, Lausanne. A.A. Balkema, Rotterdam, tional Journal of Surface Mining, Reclamation and Envi-
v.1, 531-536. ronment 13(1): 11-18.
Benko, B. & Stead, D. 1998. The Frank slide: A reexamination Crandell D.R. & Mullineaux, D.R. 1967. Volcanic Hazards at
of the failure mechanism. Canadian Geotechnical Journal Mount Rainier, Washington. USGS Bulletin 1238, 26p.
35(2): 299-311. Crandell, D.R., Mullineaux, D.R. & Miller, C.D. 1979. Vol-
Bhandari, R.K. 1988. Special lecture: Some practical lessons in canic-hazards studies in the Cascade Range of the western
the investigation and field monitoring of landslides. In United States. In Sheets & Grayson (eds.), Volcanic Activ-
Bonnard (ed.), Proceedings of the Fifth International Sym- ity and Human Ecology. Academic Press, pp. 195-219.
posium on Landslides, Lausanne. A.A. Balkema, Rotter- Crosta, G.B. & Agliardi, F. 2003. Failure forecast for large
dam, v.3, 1435-1457. rock slides by surface displacement measurements. Cana-
Bhasin, R. & Kaynia, A.M. 2004. Static and dynamic simula- dian Geotechnical Journal 40(1): 176-191.
tion of a 700-m high rock slope in western Norway. Engi- Crosta, G. B., Imposimato, S., Roddeman, D., Chiesa S. &
neering Geology 71(3-4): 213-226. Moia, F. 2005. Small fast moving flow-like landslides in
Board, M., Chacon, E., Varona, P. & Lorig, L. 1996. Compara- volcanic deposits: the 2001 Las Colinas Landslide (El Sal-
tive analysis of toppling behaviour at Chuquicamata open- vador). Geotechnique: In press.
pit mine, Chile. Transactions of the Institution of Mining Cruden, D.M. & Varnes D.J. 1996. Landslide types and proc-
and Metallurgy - Section A: Mining Industry 105: A11- esses. In Turner & Schuster (eds.), Landslide Investigation
A21. and Mitigation. TRB Special Report 247. National Acad-
Bonzanigo, L., Eberhardt, E. & Loew, S. 2001. Hydromechani- emy Press, Washington D.C. pp. 36-75.
cal factors controlling the creeping Campo Vallemaggia Davies, T.R.H. 1982. Spreading of rock avalanche debris by
landslide. In Kühne et al. (eds.), UEF International Confer- mechanical fluidization. Rock Mechanics 15: 9-24.
ence on Landslides - Causes, Impacts and Countermea- Davies T.R., McSaveney, M.J. & Hodgson, K.A. 1999. A
sures, Davos. Verlag Glückauf GmbH, Essen, 13-22. fragmentation-spreading model for long-runout rock ava-
Campbell, C.S. 1989. Self-lubrication for long runout land- lanches. Canadian Geotechnical Journal 36: 1096-1110.
slides. Journal of Geology 97: 653-665. Dawson, E.M., Roth, W.H. & Drescher, A. 1999. Slope stabil-
Casagrande, A. 1976. Liquefaction and cyclic deformation of ity analysis by strength reduction. Geotechnique 49(6):
sands: a critical review. Harvard Soil Mechanics Series, No. 835-840.
88, Cambridge, Massachussetts, 26 p. Denlinger, R.P. & Iverson, R.M. 2001. Flow of variably fluid-
Chen, H. & Lee, C.F. 2000. Numerical simulation of debris ized granular masses across three-dimensional terrain. 1.
flows. Canadian Geotechnical Journal 37:146-160. Numerical predictions and experimental tests. Journal of
Chiriotti, E., Mahtab, A., Xu, S. & Kendorski, F. 1999. Meth- Geophysical Research 106:553-566.
odology for characterizing very large rock masses: An illus- Domaas, U. 1994. Geometrical methods of calculating rockfall
trative example. In Amadei et al. (eds.), Rock Mechanics range. Norwegian Geotechnical Institute. Report 585910-1.
for Industry, Proceedings of the 37th U.S. Rock Mechanics 21 pp. Unpublished.
Symposium, Vail. A.A. Balkema, Rotterdam, v.1, 263-270. Eberhardt, E., Bonzanigo, L. & Loew, S. 2005b. Long-term in-
Chryssanthakis, P. & Grimstad, E. 1996. Landslide simulation vestigation of a deep-seated creeping landslide in crystal-
by using the distinct element method. In Aubertin et al. line rock - Part 2: Mitigation measures and numerical mod-
(eds.), Rock Mechanics: Tools and Techniques, Proceed- elling of deep drainage at Campo Vallemaggia. Canadian
ings of the 2nd North American Rock Mechanics Sympo- Geotechnical Journal: In preparation.
sium (NARMS'96), Montréal. A.A. Balkema, Rotterdam, Eberhardt, E., Spillmann, T., Maurer, H., Willenberg, H.,
v.2, 1921-1928. Loew, S. & Stead, D. 2004b. The Randa Rockslide Labora-
Close, U. & McCormick, E. 1922. Where the mountains tory: Establishing brittle and ductile instability mechanisms
walked. National Geographic Magazine 41(5): 445-464. using numerical modelling and microseismicity. In Lacerda
Coggan, J.S., Stead, D. & Eyre, J.M. 1998. Evaluation of tech- et al. (eds.), Proceedings of the 9th International Sympo-
niques for quarry slope stability assessment. Transactions sium on Landslides, Rio de Janeiro. A.A. Balkema, Leiden,
of the Institution of Mining and Metallurgy, Section B 107: v.1, 481-487.
B139-B147. Eberhardt, E., Stead, D., Coggan, J. & Willenberg, H. 2002. An
Collins, B.D. & Znidarcic, D. 2004. Stability analyses of rain- integrated numerical analysis approach to the Randa rock-
fall induced landslides. Journal of Geotechnical and slide. In Rybár et al. (eds.), Proceedings of the 1st European
Geoenvironmental Engineering 130(4): 362-372. Conference on Landslides, Prague. A.A. Balkema, Lisse,
Copons, R. 2004. Avaluació de la perillositat de caiguda de 355-362.
blocs a Andorra la Vella (Principat d’Andorra). PhD The- Eberhardt, E., Stead, D. & Coggan, J.S. 2004a. Numerical
sis. Universitat de Barcelona. Unpublished. 244 pp. analysis of initiation and progressive failure in natural rock
Corominas, J. 1996. The angle of reach as a mobility index for slopes - the 1991 Randa rockslide. International Journal of
small and large landslides. Canadian Geotechnical Journal Rock Mechanics and Mining Sciences 41(1): 69-87.
33: 260-271. Eberhardt, E., Thuro, K. & Luginbuehl, M. 2005a. Slope insta-
Corominas, J., Esgleas, J. and Baeza, C. 1990. Risk mapping in bility mechanisms in dipping interbedded conglomerates
the Pyrenees area: a case study. Hydrology in Mountainous and weathered marls - the 1999 Rufi landslide, Switzerland.
Regions II. IAHS Publication 194: 425-428 Engineering Geology 77(1-2): 35-56.
Corominas, J., Copons, R., Vilaplana, J.M., Altimir, J. & Evans, S.G. & Hungr, O. 1993. The assessment of rockfall haz-
Amigó, J. 2003b. Integrated landslide susceptibility analy- ard at the base of talus slopes. Canadian Geotechnical
sis and hazard assessment in the Principality of Andorra. Journal 30: 620-636.
Natural Hazards 30: 421-435. Faella, C. & Nigro, E. 2001. Effetti delle colate rapide sulle
costruzioni. Parte Prima: Descrizione del danno. Forum per
il Rischio Idrogeologico ‘‘Fenomeni di colata rapida di
fango nel Maggio ’98’’, Naples, June 22 giugno 2001:105– Hungr, O. 2003. Flow slides and flows in granular soils. Key-
112. note Paper. In Picarelli (ed.), Proc., FLOWS 2003, Interna-
Fannin, R.J. & Wise, M.P. 2001. An empirical-statistical model tional Workshop, Sorrento, Italy, Kluwer Publishers.
for debris flow travel distance. Canadian Geotechnical Hungr, O. 2004. Prospects for prediction of landslide dam ge-
Journal 38: 982-994. ometry using dynamic models. Proceedings, NATO confer-
Federico, A., Fidelibus, C. & Interno, G. 2002. The prediction ence on Landslide dams, Bishkek, Kyrghistan, July, 2004,
of landslide time to failure - A state of the art. In Proceed- In review.
ings of the 3rd International Conference on Landslides, Hungr, O. 2005. Rock avalanche occurrence, process and mod-
Slope Stability and the Safety of Infra-Structures, Singa- elling. Keynote Paper, NATO Advanced Workshop on
pore, 167-180. Massive Slope Failure, Celano, Italy. Kluwer NATO Sci-
Finlay, P.J., Mostyn, G.R. & Fell, R. 1999. Landslide risk as- ence Series, In press.
sess-ment: prediction of travel distance. Canadian Geo- Hungr, O., Dawson, R., Kent, A., Campbell, D. & Morgen-
technical Journal 36: 556-562. stern, N.R. 2002. Rapid flow slides of coal mine waste in
Fletcher, L., Hungr, O. & Evans, S.G. 2002. Contrasting failure British Columbia, Canada. In Catastrophic Landslides Geo-
behaviour of two large landslides in clay and silt. Canadian logical Society of America Reviews in Engineering Geol-
Geotechnical Journal 39:46-62. ogy No. 14.
Fredlund, D.G. & Krahn, J. 1977. Comparison of slope stabil- Hungr, O. & Evans, S.G. 1988. Engineering evaluation of
ity methods of analysis. Canadian Geotechnical Journal fragmental rockfall hazards. In Bonnard (ed.), 5th Interna-
14(3): 429-439. tional Symposium on Landslides, Lausanne, Switzerland.
Fukuzono, T. 1990. Recent studies on time prediction of slope A.A. Balkema, vol. 1: 685-690.
failure. Landslide News 4: 9-12. Hungr, O. & Evans, S.G. 1996. Rock avalanche runout predic-
García, J.L., López, J.L., Noya, M., Bello, M.E., Bello, M.T., tion using a dynamic model. Proceedings, 7th International
Gonzájez, N., Paredes, G. & Vivas, M.I. 2003. Hazard Symposium on Landslides, Trondheim, Norway, 1: 233-
mapping for debris flow events in the alluvial fans of north- 238.
ern Venezuela. In Rickenmann & Chen (eds.), Debris Flow Hungr, O. & Evans, S.G. 2004a. Entrainment of debris in rock
Hazards Mitigation: Mechanics, Prediction and Assess- avalanches; an analysis of a long run-out mechanism. Bulle-
ment. Millpress, Rotterdam, 1:589-600. tin, Geological Society of America, In press.
Geertsema, M., Hungr, O., Schwab, J.W. & Evans, S.G. 2004. Hungr, O. & Evans, S.G. 2004b. The occurrence and classifica-
A large rock slide – debris avalanche at Pink Mountain, tion of massive rock slope failure. Felsbau 22: 16-23.
northeastern British Columbia, Canada. Engineering Geol- Hungr, O., Evans, S.G., Bovis, M. & Hutchinson, J.N. 2001.
ogy: In press. Review of the classification of landslides of the flow type.
Gray, J.M.N.T., Wieland, M. & Hutter, K. 1999. Gravity Environmental and Engineering Geoscience VII: 221-238.
driven free surface flow of granular avalanches over com- Hunter, G. & Fell, R. 2003. Travel distance angle for rapid-
plex basal topography. Proceedings, Royal Society of Lon- landslides in constructed and natural soil slopes. Canadian
don, Ser. A 455: 1841-1874. Geotechnical Journal 40(6): 1123-1141.
Griffiths, D.V. & Lane, P.A. 1999. Slope stability analysis by Hutchinson, J.N. 1988. Morphological and geotechnical pa-
finite elements. Geotechnique 49(3): 387-403. rameters of landslides in relation to geology and hydro-
Gruner, U. 2001. Felssturzgefahr Chapf - Üssri Urweid (Ge- geology. In Bonnard (ed.), 5th International Symposium on
meinde Innertkirchen): Angaben zur Geologie und zu den Landslides, Lausanne, Switzerland. A.A. Balkema, vol. 1:
Bewegungen. Bern, Kellerhaus + Haefeli AG, 6 pp. 3-35
Gruner, U. 2003. Zweite Sicherheitssprengung am Chapf bei Hungr, O. & Kent, A. 1995. Coal mine waste dump failures in
Innertkirchen vom 20. August 2002. Bulletin für Ange- British Columbia, Canada. Landslide News 9: 26-28.
wandte Geologie 8(1): 17-25. Hungr, O. & Morgenstern, N.R. 1984. Experiments in high ve-
Hajiabdolmajid, V. & Kaiser, P.K. 2002. Modelling slopes in locity open channel flow of granular materials. Géotech-
brittle rock. In Hammah et al. (eds.), Proceedings of the 5th nique 34: 405-413. Discussion and Reply, 35: 383-385.
Northern American Rock Mechanics Symposium Hutchinson, J.N. 1986. A sliding-consolidation model for flow
(NARMS-TAC), Toronto. University of Toronto Press, To- slides. Canadian Geotechnical Journal 23: 115-126.
ronto, v.1, 331-338. Hutchinson, J.N., Prior, D.B. & Stephens, N. 1974, Potentially
Hart, R.D. 1993. An introduction to distinct element modelling dangerous surges in an Antrim mudslide. Quarterly Journal
for rock engineering. In Hudson (ed.), Comprehensive of Engineering Geology 7: 363-376.
Rock Engineering: Principles, Practice & Projects. Oxford: Itasca 2000. UDEC - Universal Distinct Element Code. Itasca
Pergamon Press, v.2, 245-261. Consulting Group, Inc., Minneapolis.
Heim, A. 1932. Bergsturz und Menschenleben. Zurich: Fretz Iverson, R.M. 1997. The physics of debris flows. Reviews of
and Wasmuth Verlag, 218 pp. Geophysics 35(3): 245-296.
Hoblitt, R.P., Walder, J.S., Driedger, C.L., Scott, K.M., Iverson, R.M. 2005. Debris flow mechanics. Chapter 7 in Ja-
Pringle, P.T. & Vallance, J.W. 1998. Volcano Hazards from kob & Hungr (eds.), Debris Flows and Related Phenomena.
Mount Rainier, Washington, Revised 1998. USGS Open- Praxis, Springer, Heidelberg, In press.
File Report 98-428 Iverson, R.M. & Denlinger, R.P. 2001. Flow of variably fluid-
Hsü, K.J. 1975. Catastrophic debris stream (sturzstroms) gen- ized granular masses across three-dimensional terrain. 1.
erated by rockfalls. Geological Society of America Bulletin Coulomb mixture theory. Journal of Geophysical Research
86: 129-140. 106: 537-552.
Hungr, O. 1990. Mobility of rock avalanches. Report of the Iverson, R.M., Logan, M. & Denlinger, R.P. 2004. Granular
National Institute for Earth Science and Disaster Preven- avalanches across irregular three-dimensional terrain: 2.
tion, 46: 11-20 Experimental tests. Journal of Geophysical Research 109:
Hungr, O. 1995. A model for the runout analysis of rapid flow 16 pp.
slides, debris flows and avalanches. Canadian Geotechnical Jackson, L.E., Kostashuk, R.A. & MacDonald, G.M. 1987.
Journal 32: 610-623. Identification of debris flow hazard on alluvial fans in the
Hungr, O. 2000. Analysis of debris flow surges using the the- Canadian Rocky mountains. Geological Society of America.
ory of uniformly progressive flow. Earth Surface Processes Reviews in Engineering Geology VII: 155-124.
and Landforms 25: 1-13.
Jeyapalan, K. 1981. Analysis of flow failures of mine tailings posium on Landslides, Lausanne. A.A. Balkema, Rotter-
impoundments. Ph.D. Thesis, University of California, dam, v.1, 757-762.
Berkeley. Sassa, K. 1985. The mechanism of debris flows. Proceedings,
Johnson, A.M. 1970. Physical Processes in Geology. W.H. XI International Conference on Soil Mechanics and Foun-
Freeman, New York, 577pp. dation Engineering, San Francisco, Vol. 1, pp. 1173-1176.
Kennedy, B.A. & Niermeyer, K.E. 1970. Slope monitoring sys- Sassa, K. 1988. Geotechnical model for the motion of land-
tems used in the prediction of a major slope failure at the slides. In Bonnard (ed.), 5th International Symposium on
Chuquicamata Mine, Chile. In Van Rensburg (ed.), Plan- Landslides, Lausanne, Switzerland. A.A. Balkema, vol. 1:
ning Open Pit Mines, Proceedings, Johannesburg. A.A. 37-55.
Balkema, Cape Town, 215-225. Sassa, K., 2002. Mechanism of rapid and long traveling flow
Körner, H.J. 1976. Reichweite und Geschwindigkeit von phenomena in granular soils. Proceedings, International
Bergsturzen und fleisschnee-lawinen. Rock Mechanics 8: Symposium on Landslide Mitigation and Protection of Cul-
225-256. tural and Natural Heritage. Kyoto University, Kyoto, Japan,
Li, T. 1983. A mathematical model for predicting the extent of 11-30.
a major rockfall. Zeitschrift für Geomorphologie 24: 473- Savage, S.B. & Hutter, K. 1989. The motion of a finite mass of
482. granular material down a rough incline. Journal of Fluid
Lo, K.Y. & Lee, C.F. 1973. Stress analysis and slope stability Mechanics 199: 177-215.
in strain-softening materials. Geotechnique 23(1): 1-11. Scheidegger, A. 1973. On the prediction of the reach and ve-
Löw, S. 1997. Wie sicher sind geologische Prognosen? Bulletin locity of catastrophic landslides. Rock Mechanics 5: 231-
für Angewandte Geologie 2(2): 83-97. 236.
Main, I.G., Sammonds, P.R. & Meredith, P.G. 1993. Applica- Scheller, E. 1971. Beitrag zum bewegungsverhalten grosser
tion of a modified Griffith criterion to the evolution of frac- bergstürze. Eclogae Geologicae Helvetiae 64: 195-202.
tal damage during compressional rock failure. Geophysical Scott, K.M. & Vallance, J.W. 1993. History of Landslides and
Journal International 115(2): 367-380. Debris Flows at Mount Rainier: Water Fact Sheet. USGS
McDougall, S. & Hungr, O. 2004a. A model for the analysis of Open-File Report 93-111.
rapid landslide runout motion across three-dimensional ter- Segre, A. & DeAngeli, C. 1995. Cellular automaton for realis-
rain. Canadian Geotechnical Journal 41(6): 1084-1097. tic modelling of landslides. European Geophysical Society,
McDougall, S. & Hungr, O. 2004b. A universal model for 3-D Non-linear Processes in Geophysics.
runout analysis of landslides. Proceedings, 9th International Shou, K.-J. & Wang, C.-F. 2003. Analysis of the Chiufenger-
Symposium on Landslides, Rio de Janeiro, Brazil. shan landslide triggered by the 1999 Chi-Chi earthquake in
McDougall, S. & Hungr, O. 2005. A dynamic model for rapid Taiwan. Engineering Geology 68(3-4): 237–250.
landslides that entrain material from the path. Canadian Shreve, R.L. 1968. The Blackhawk landslide. Geological Soci-
Geotechnical Journal: In review. ety of America, Special Paper 108.
Michael-Leiba, M., Baynes, F., Scott, G. & Granger, K. 2003. Soussa, J. & Voight, B. 1991. Continuum simulation of flow
Regional landslide risk to the Cairns community. Natural failures. Geotechnique 41: 515-538.
Hazards 30: 233-249. Stead, D. & Coggan, J.S. 2005. Numerical modelling of rock
Munjiza, A., Owen, D.R.J. & Bicanic, N. 1995. A combined slopes using a total slope failure approach. In Evans et al.
finite-discrete element method in transient dynamics of (eds.), Massive Rock Slope Failure: New Models for Haz-
fracturing solids. Engineering Computations 12: 145-174. ard Assessment, Proceedings of the NATO Advanced Re-
Nichol, S.L., Hungr, O. & Evans, S.G. 2002. Large scale brittle search Workshop, Celano, Italy. Kluwer Academic Pub-
and ductile toppling of rock slopes. Canadian Geotechnical lishers, Dordrecht, In Press.
Journal 39(4): 773-788. Stead, D. & Eberhardt, E. 1997. Developments in the analysis
Nicoletti, P.G. & Sorriso-Valvo, M. 1991. Geomorphic con- of footwall slopes in surface coal mining. Engineering Ge-
trols of the shape and mobility of rock avalanches. Geo- ology 46(1): 41-61.
logical Society of America Bulletin 103: 1365-1373. Stead, D., Eberhardt, E., Coggan, J. & Benko, B. 2001. Ad-
O’Brien, J.S., Julien, P.Y. & Fullerton, W.T. 1993. Two- vanced numerical techniques in rock slope stability analysis
dimensional water flood and mudflow simulation. Journal - Applications and limitations. In Kühne et al. (eds.), Inter-
of the Hydraulics Division, ASCE 119(HY2): 244-261. national Conference on Landslides - Causes, Impacts and
Okura, Y., Kitahara, H., Sammori, T. & Kawanami, A. 2000. Countermeasures, Davos. Verlag Glückauf Essen, Essen,
The effects of rockfall volume on runout distance. Engi- 615-624.
neering Geology 58: 109-124. Suwa, H. 1991. Visually observed failure of a rock slope in Ja-
Petley, D.N., Bulmer, M.H. & Murphy, W. 2002. Patterns of pan. Landslide News 5: 8-10.
movement in rotational and translational landslides. Geol- Takahashi, T. 1991. Debris flow. IAHR Monograph, A.A.
ogy 30(8): 719-722. Balkema, Rotterdam, 165 pp.
Pitman, E.B., Nichita, E.B., Patra, A.K., Bauer, A.C., Bursik Terzaghi, K. 1950. Mechanism of landslides. In Paige (ed.),
M. & Webb, A. 2003. A model of granular flows over an Application of Geology to Engineering Practice (Berkey
erodible surface. Discrete and Continuous Dynamical Sys- Volume). New York: Geological Society of America, 83-
tems–Series B 3: 589–599. 123.
Revellino, P., Hungr, O., Guadagno, F.M. & Evans, S.G. 2003. Trunk, F.J., Dent, J.D. & Lang, T.E. 1986. Computer model-
Velocity and runout prediction of destructive debris flows ling of large rockslides. Journal of the Geotechnical Divi-
and debris avalanches in pyroclastic deposits, Campania sion, ASCE 112(GE3): 348-360.
Region, Italy. Environmental Geology 45:295-311. Van Gassen, W. & Cruden, D. 1989. Momentum transfer and
Saito, M. 1965. Forecasting the time of occurrence of a slope friction in the debris rock avalanches. Canadian Geotechni-
failure. In Proceedings of the Sixth International Confer- cal Journal 26: 623-628.
ence on Soil Mechanics and Foundation Engineering, Varnes, D.J. 1978. Slope movement types and processes. In
Montreal. University of Toronto Press, Toronto, v.2, 537- Schuster & Krizek (eds.), Landslides, Analysis and Control:
541. Transportation Research Board, National Academy of Sci-
Salt, G. 1988. Landslide mobility and remedial measures. In ences, Washington, DC., Special Report 176: 11-33.
Bonnard (ed.), Proceedings of the Fifth International Sym- Voellmy, A. 1955. Über die Zerstorungskraft von Lawinen.
Schweiz. Bauzeitung 73: 212-285.
Voight, B. 1978. The Lower Gros Ventre slide, Wyoming,
U.S.A. In Voight (ed.). Rockslides and avalanches. El-
sevier, Amsterdam. vol. 1: 113-166.
Voight, B. 1989. A relation to describe rate-dependent material
failure. Science 243: 200-203.
Voight, B. & Kennedy, B.A. 1979. Slope failure of 1967-1969,
Chuquicamata Mine, Chile. In Voight (ed.), Rockslides and
Avalanches. Amsterdam: Elsevier Scientific Publishing
Company, 595-632.
Voight, B. & Sousa, J. 1994. Lessons from Ontakesan: A com-
parative analysis of debris avalanche dynamics. Engineer-
ing Geology 38: 261–297.
Willenberg, H., Spillmann, T., Eberhardt, E., Evans, K., Loew,
S. & Maurer, H. 2002. Multidisciplinary monitoring of pro-
gressive failure processes in brittle rock slopes - Concepts
and system design. In Rybár et al. (eds.), Proceedings of the
1st European Conference on Landslides, Prague. A.A.
Balkema, Lisse, 477-483.
Zienkiewicz, O.C. & Huang, M. 1995. Localization problems
in plasticity using finite elements with adaptive remeshing.
International Journal for Numerical and Analytical Meth-
ods in Geomechanics 19(2): 127-148.
Zvelebil, J. 1984. Time prediction of a rockfall from a sand-
stone rock slope. In Proceedings of the Fourth International
Symposium on Landslides, Toronto. University of Toronto
Press, Downsfield, v.3, 93-95.

You might also like