You are on page 1of 14

Original Article

Landslides (2005) 2: 8396


DOI: 10.1007/s10346-005-0049-1
Received: 9 October 2004
Accepted: 7 March 2005
Published online: 12 May 2005
Springer-Verlag 2005

Jordi Corominas Jose Moya Alberto Ledesma Antonio Lloret Josep A. Gili

Prediction of ground displacements and velocities


from groundwater level changes at the Vallcebre landslide
(Eastern Pyrenees, Spain)

Abstract In active landslides, the prediction of acceleration of movement is a crucial issue for the design and performance of warning
systems. The landslide of Vallcebre in the Eastern Pyreenes, Spain,
has been monitored since 1996 and data on rainfall, groundwater
levels and ground displacements are measured on a regular basis.
Displacements observed in borehole wire extensometers have shown
an immediate response of the landslide to rainfall episodes. This rapid
response is likely due to the presence of preferential drainage ways.
The occurrence of nearly constant rates of displacement in coincidence with steady groundwater levels suggests the presence of viscous
forces developed during the movement. An attempt to predict both
landslide displacements and velocities was performed at Vallcebre by
solving the momentum equation in which a viscous term (Bingham
and power law) was added. Results show that, using similar rheological parameters for the entire landslide, computed displacements
reproduce quite accurately the displacements observed at three selected wire extensometers. These results indicate that prediction of
displacements from groundwater level changes is feasible.
Keywords Translational slide . Viscous behaviour . Prediction
of displacements . Eastern Pyrenees . Spain
Introduction
Traditional strategies in landslide hazard management have been
mostly oriented at avoiding dangerous sites and stabilizing unstable slopes. Regional and urban planning can be powerful tools to
ensure that development occurs in the safest places. In that respect,
planners and decision-makers have benefited from the information
provided by hazard and risk maps. However, development in many
regions has occurred, and indeed continues to occur without hazard
maps being available. Because of this, there are numerous examples
of development built either on or near large landslides. In such cases,
remedial measures are often unaffordable, while moving the population to more stable slopes can pose a considerable societal problem.
As a consequence, the coexistence of human activities with natural
hazards must be considered with the risk being mitigated through
the use of hazard management strategies.
In landslide threatened areas, solutions usually focus on the implementation of mitigation measures that avoid the slope failure or
divert the moving mass away from vulnerable elements, or protect
or reinforce the threatened elements. However, in some places, to
stabilise a landslide may simply be too costly in financial and/or environmental terms. When mitigation is not feasible, it is important
to consider the implementation of warning systems that may, at least,
avoid damage and/or loss of human lives. To be effective and reliable, warning systems require a sound knowledge of the behaviour
of the landslide, including the mechanism, the potential triggers and
their respective thresholds (i.e. rainfall intensity and duration), the

expected time of failure or reactivation, the expected velocity, and the


areal extent including the runout distance where appropriate.
There are many uncertainties in the forecasting of when a movement in a landslide will occur, or when rates of movement will increase
and thus it can not be performed reliably at present. Therefore, improving this capability is a basic requirement for the development of
warning systems.
The existing methods of predicting and forecasting the landslide
occurrence have mostly concentrated on first-time slope failures. The
simplest method for predicting landslide occurrence at a regional
scale is through the establishment of empirical rainfall thresholds.
This method is based on the analysis of storms that have produced
slope failures. The fundamental assumption is that, in a homogeneous
region, landslides occur once a precipitation threshold has been overcome. The threshold is usually expressed as rainfall intensity-duration
relationship (Caine 1980; Cancelli and Nova 1985; Wieczorek 1987;
Larsen and Simon 1993) and it has been used with some degree of
success as a warning system in areas affected by widespread and recurrent shallow landsliding (Keefer et al. 1987).
A more precise prediction of the time of failure can be obtained by
measuring the displacement rates in slopes and landslides. Translational and rotational movements display a range of behaviour styles,
including long-term creep, catastrophic movement that is preceded
by long-term creep, and sudden catastrophic movement with no creep
phase (Petley and Allison 1997). Forecasting of catastrophic event may
be made by analyzing the late stages before the slope failure, i.e. tertiary creep (Saito 1965). Time of failure can be determined by curve
fitting techniques (Voight and Kennedy 1979; Picarelli et al. 2000), or
from the linear trend of the inverse rate curve (Voight 1989; Fukuzono
1990). A summary of the different methods proposed in the literature
can be found in Federico et al. (2004).
Prediction of the failure may be also carried out by physically-based
numerical models. Significant developments have been produced
over the last years in slope stability models (Matsui and San 1992;
Griffiths and Lane 1999) and, in particular, in coupled hydrological
and mechanical models (Wilkinson et al. 2002). Numerical models
can provide a good understanding of the mechanism of failure and
they can consider complex landslide geometries, spatial variations
in soil properties and three-dimensional groundwater flow, among
other advantages. However, the applicability of these models to the
prediction of slope displacements and failure conditions is still far
from routine, is complex and time consuming, and experiences of 3D
coupled modelling are rather limited.
Suspended and dormant landslides can be reactivated in periods
of heavy rainfall while active ones may show phases of acceleration
and deceleration. This variable movement pattern has been observed
in slow-moving earthflows (Lateltin and Bonnard 1995), mudslides
(Angeli et al. 1996), and in both rotational and translational landslides

Landslides 2 . 2005

83

Original Article
(Corominas et al. 1999). In some of these movements, the search for
reliable rainfall thresholds has been unsuccessful (Noverraz et al. 1998;
Flageollet et al. 1999) and alternative approaches are required to assess
the conditions leading to the reactivation. Petley et al. (2002) analyzed
inverse rate curves and demonstrated that landslides occurring in
existing failure surfaces show an increase of the rate of deformation
up to a constant rate (steady-state behaviour) for any given stress state
material combination. On the other hand, the use of hydrological
and mechanical models is restricted because of the complexity of
these landslides. Existing models remain limited when considering
groundwater changes and rates of displacement. Some attempts to
examine large landslides have included the use of simplified empirical
models that combine hydrological and stability analyses (Van Asch
and Buma 1997) and the use of viscous constitutive laws for the
simulation of continuous displacement patterns (Vulliet 2000).
Continuously moving landslides require a dynamic analysis instead of a classical static approach. The temporal scale is fundamental, and any modelling attempt should consider this. In addition to
these conceptual difficulties, the analysis of the response to both short
and medium-term rainfall scenarios with hydrological and mechanical models requires long records of high quality data that include
rainfall, groundwater levels and landslide displacements.
In this paper we present an examination of movement of the Vallcebre landslide, a slow-moving landslide activated by rainfall. A simulation of the displacement pattern in relation to groundwater record
by means of a dynamic, physical model has been attempted. The
work has been complemented by an extensive field survey and laboratory investigation to allow the best possible modelling of landslide
behaviour. That has included an analysis of the rheology of the sliding material and of the influence of the viscous component on the
motion.

The Vallcebre landslide


The Vallcebre landslide is a large, active slope failure located in the
upper Llobregat river basin, in the Eastern Pyrenees, 140 km north
of Barcelona, Spain. The landslide is situated on the western slope of
the Serra de la Llacuna (Fig. 1). The mobilised material consists of a
set of shale, gypsum and claystone layers of continental origin gliding
over a thick limestone bed, all of which are of Upper Cretaceous
Lower Palaeocene age. The dimensions of the slide mass are 1200 m
long and 600 m wide. The entire landslide involves an area of 0.8 km2
that shows superficial cracking and distinct ground displacements.
The age of the landslide is not known but it is known to have been
active for several centuries at least.
From a geomorphological point of view, the Vallcebre landslide is
of a translational type (Fig. 2). A longitudinal profile shows a stairshape with three main slide units of decreasing thickness towards the
landslide toe (Fig. 3). Each unit is formed by a gentle slope surface
bounded in its downhill edge by a scarp of a few tens of meters high.
At the base of each scarp an extension area develops in the form
of a graben. This fact is interpreted as the lower units moving more
rapidly than upper ones, which has been confirmed by the monitoring
network installed. The average slope of the landslide is about 10 .
The toe of the landslide extends to the Vallcebre torrent bed, and is
pushing it towards the opposite bank. As a result of this, the Vallcebre
torrent has been shifted to the west more than ten meters and the
foot of the landslide has overridden the opposite slope to form a
back tilted surface (Fig. 4). The torrent undermines the landslide toe
during floods, causing local rotational failures which decrease the
overall stability.
Most of the evidence of surface deformation is situated at the
boundaries of the slide units in the form of distinct shear surfaces
and tension cracks. At the base of each transverse scarp, both the

Fig. 1 General view of the Vallcebre translational slide. The movement is bounded, in the background, by the high scarp of Serra de La Llacuna (center-right), in the foreground by the Vallcebre
torrent, which runs from the right to the left in an incised gully, and laterally by the light grey limestone outcrops

84

Landslides 2 . 2005

Fig. 2 A geomorphological sketch of the Vallcebre landslide

Fig. 3 Geological cross-sections of the Vallcebre landslide (the location of the profiles appears in Fig. 2)

Landslides 2 . 2005

85

Original Article

Fig. 4 Photograph of the toe of the Vallcebre landslide which is being continuously undermined and eroded by the Vallcebre torrent. Local slope failures are observable in the front

ground surface and the trees are tilted backwards due to the development of a graben along with a slight rotation of the head. In
contrast, within the units, the ground surface is only disturbed by
minor fissures, scarps less than 50 m long, and by some cracking
of the walls of farm-houses standing on the landslide. The direction of both the transverse scarps and grabens, indicates a movement towards the north-west (see Fig. 2). A secondary direction of
movement, towards the Torrent Llarg, is also suggested by the trend
of the escarpment of the upper slide unit, but this has not been
validated to date. The most active area is the lower unit, which is
bounded, at the south-western side, by the torrents of Vallcebre and
Llarg and, at its north-eastern side, by a well developed lateral shear
surface.
The geological structure of the landslide has been obtained by
means of an intense geological and geomorphological assessment,
which has included mapping of the surficial exposures, geophysical
surveys and drilling. The mobilised material consists of a sequence
of continental sediments. From the bottom to the top it includes
(Fig. 3): (i) densely fissured shales, 1 to 6 m thick, showing distinct
slickensides; (ii) locally, gypsum lenses up to 5 m thick and some tens
of meters width which are predominant in the southern side of the
landslide; and (iii) clayey siltstones rich in veins and micronodules of
gypsum with thickness of up to tens of meters. In addition to these
layers, the extension zones located at the toe of the scarps have been
filled with colluvium composed of boulders and gravel with a silty
matrix.
Gypsum lenses are affected by solution processes and have developed some karst features like pipes and springs. These pipes run

86

Landslides 2 . 2005

close to the topographical surface and frequent collapses (sinkholes)


are observable in the ground surface, which are arranged following
approximately straight paths.
The area is affected by two thrust faults and some associated folds.
One of these faults runs underneath the landslide and has a vertical
offset of about 10 m (Fig. 3). The landslide also overlays and hides a
tightly folded syncline that forms a thick core of red clayey siltstones.
This syncline has an east-west trend and its axis dips towards the west,
more or less parallel to the down-slope direction.

Monitoring of the landslide


The lower unit of the Vallcebre landslide has been monitored since
1987 using conventional surveying and photogrammetry (Gili and
Corominas 1992). During July 1996, March 1997 and April 1998, sixteen boreholes were drilled in the landslide in order to: log the geologic materials of the landslide; to provide undisturbed samples for
laboratory tests; to allow in-situ hydrological testing; and to set up
a monitoring network. Boreholes were equipped with inclinometers,
wire extensometers and piezometers. Further details about the site,
the fieldwork carried out and the monitoring scheme can be found
in Corominas et al. (1999, 2000).
The failure surface of the landslide was determined using inclinometers. The inclinometric profiles showed that the displacement
occurs in a thin basal shear zone, with negligible deformation above
it (Corominas et al. 1999). The shear zone develops through the fissured clayey siltstone layer, close to the contact with the limestone. It
has an average inclination of 10 towards the Vallcebre torrent and it

Fig. 5 Wire displacements at boreholes S-2, S-5, S6, S9 and S-11

runs roughly parallel to the ground surface except in the area close
to the contact between the intermediate and lower units where the
presence of the inverse fault produces a bedrock threshold and a decrease of the landslide thickness. The depth of the failure surface is
not constant. Inclinometric readings showed that the lower slide unit
has a thickness of 10 to 15 m, whereas the intermediate unit reaches
a thickness of at least 34 m in the northern side and between 14 and
19 m in the southern one.
Since 1996, systematic logging of rainfall, groundwater level
changes, and landslide displacements has been carried out every
20 min. Piezometric readings have indicated that changes in groundwater levels occur quickly. The extensometer has recorded sudden
changes in displacement rates that can be directly related to the fluctuations of the water table governed by rainfall.
Measurements of surface displacements using differential GPS have
been used to complement the measurements of the inclinometers
and wire extensometers. A total of 30 points were positioned on the
landslide surface for periodic control. These points included reference

points, stable points adjacent of the landslide, and targets within the
landslide mass (buildings, outcropping rock blocks, steel rods and
upper ends of the boreholes). Real Time Kinematics and Fast Static
were used to analyze the GPS observations (Gili et al. 2000). Fourteen
GPS campaigns were carried out from December 1995 to February
1998. During this period horizontal displacements up to 1.6 m and
subsidence of the landslide surface of up to 0.35 m were observed.
The wire extensometer measurements show that the landslide has
never completely stopped moving since we started the continuous
monitoring in November 1996, although velocities reduced significantly during dry periods (Fig. 5). On the other hand, the history
of displacement of the extensometers reflects that different parts of
the landslide mass move synchronically but with a different rate of
displacement. Extensometer S2 has shown the fastest displacements,
with maximum recorded rates of up to 50 mm/week. The other extensometers standing on this slide unit, the S5, S9 and S11, have exhibited
rates lower than those of S2, although they were of the same order of
magnitude.
At borehole S6, placed on the intermediate slide unit, the velocity
is significantly smaller, indicating that, compared to the lower unit,
the intermediate unit is less active.
Determination of the landslide parameters
Tests on undisturbed samples obtained by drilling were carried out
to determine parameters and soil characteristics to be used in the
numerical analyses. Special attention was paid to tests for determining
the shear strength of the soil involved in the slip surface.
Basic identification tests and different types of shear tests were
performed on 14 undisturbed samples. Shear tests were carried out
predominantly on samples of the fissured shale unit where the slip
surface is located; in particular, some tests were conducted on observed pre-existing shear surfaces (Fig. 6). Table 1 shows the location
where the samples were taken and the type of tests performed on each
sample.
The average Atterberg limits obtained on two samples of clayey
siltstone and seven samples of fissured shales are indicated in
Table 2.

Table 1 Location of samples and tests performed


Borehole

Depth of slip surface (m)

Depth of sample (m)

Material

Type of test and specimen


Direct shear test
Ring shear test

S-2

15

15.2

Fissured shale

S-3

9.5

S-4

9.5

S-6

44.5

19.6
10.3
10.7
4.0
8.9
44.1

Fissured shale
Fissured shale
Fissured shale
Clayey siltstone
Clayey siltstone
Fissured shale

S-7

34

S-9
S-11
S-12
S-13
S-14

14.5
8
17.5
13.5
13.5

25.4
34.2
15.2
6.5
11.2
11.4
8.7

Fissured shale
Fissured shale
Fissured shale
Clayey siltstone
Clayey siltstone
Clayey siltstone
Clayey siltstone

Existing shear
surface
Undisturbed
Undisturbed

Undisturbed
Undisturbed
Existing shear
surface
Undisturbed
Undisturbed
Undisturbed
Undisturbed
Undisturbed
Undisturbed

CU triaxial test

Remoulded

Undisturbed

Remoulded

Remoulded
Remoulded
Remoulded

Undisturbed

Remoulded

Remoulded

Remoulded
Remoulded
Remoulded

Undisturbed

Landslides 2 . 2005

87

Original Article
Fig. 6 Slickensided shear surface within
the fissured shales

Table 2 Average Atterberg limits


Fissured shales
Clayey siltstones

Liquid limit

Plastic limit

Plasticity index

41.8
54.5

22.1
35.2

19.7
19.3

Using a conventional direct shear box apparatus, a displacement,


totalling 30 mm was applied in three steps, consisting of an initial
displacement of 7.5 mm followed by a return displacement of 15 mm,
and finally a displacement of 7.5 mm to return to the starting position. A minimum of three different vertical stresses were applied on
each specimen (50 or 60 mm diameter, 25 mm high). The rate of displacement was 7.5 mm/day ensuring that shear occurred under fully
drained conditions. For two samples of fissured shale, taken from a
depth of 15.2 m from borehole S2 and from a depth of 44.1 m from

Table 3 Peak shear strength parameters of fissured shale and clayey siltstone

88

Range of normal stress (n)

c (kPa)

()

0<n<100 kPa
100<n<700 kPa

0.0
53

38.7
15.1

Landslides 2 . 2005

borehole S6, the specimen was prepared and mounted in the shear
box such that a pre-existing shear surface was the slip surface in the
shear box. In addition, direct shear tests were also performed on five
samples of fissured shale and five samples of clayey siltstone.
Ring shear tests were performed on ten samples using a Bromhead ring shear apparatus. Specimens were prepared by remoulding
the specimens previously tested in the shear box to a water content similar to their liquid limit. A total slip displacement of about
150 cm was applied in each test using displacement rates between 0.02
and 0.3 mm/min. Finally, consolidated-undrained (CU) triaxial tests

Table 4 Shear strength parameters derived from different tests : Minimum strength
measured in direct shear, ring shear and triaxial tests for fissured shale; minimum
strength measured in direct shear for clayey siltstone and residual strength measured on
a pre-existing shear surface in fissured shale
Material

Range of normal stress (n)

c (kPa)

()

Fissured shale

0<n<200 kPa
200<n<700 kPa
0<n<400 kPa
0<n<800 kPa

0.0
44.0
0.0
0.0

23.4
11.8
33.0
7.8

Clayey siltstone
Shear surface

Fig. 7 Minimum shear strength obtained


in different types of tests and samples

were performed on four specimens (50 mm diameter, 100 mm high)


extracted from three undisturbed samples of fissured shale.
Peak shear strength envelopes obtained from direct shear tests and
triaxial tests are similar for fissured shale and clayey siltstone. The
resulting failure envelopes are non linear and may be defined by the
parameters shown in Table 3.
Figure 7 shows for all three types of experiment the minimum shear
strength measured for each sample. In fissured shale the differences
between the minimum shear strength measured in direct shear, in
triaxial tests or in ring shear tests are small and are comparable to
differences between samples of the same material. The minimum
strength measured from direct shear tests on clayey siltstone is lightly
greater than the same type of strength obtained for fissured shale.
Nevertheless, the minimum strength of clayey siltstone, obtained

from ring shear tests, is similar to the minimum strength measured


for the fissured shale.
The strength measured on pre-existing shear surfaces is smaller
than the minimum strength obtained from direct shear tests and ring
shear tests, indicating that the residual state has not been reached in
these tests. Therefore, for the fissured shale, the residual strength may
be overestimated if it is assumed to be equal to the minimum strength
measured in direct shear box or in the Bromhead ring shear apparatus. Table 4 shows the parameters that define the failure envelope
associated to the minimum strength measured. Those parameters
correspond to the straight lines depicted in Fig. 7.
From volumetric changes recorded during consolidation phase
in shear box tests, the variation with applied vertical stress of the
confined elastic modulus, the consolidation coefficient, the hydraulic
conductivity, and, the secondary consolidation coefficient were determined. Typical values of confined elastic modulus are about 100 MPa
and hydraulic conductivities range between 107 and 109 cm/s.
Hydrological changes and landslide response
Groundwater variations for the period November 1996 to October
1997, measured using six piezometers (see Fig. 2 for the locations) are
shown in Fig. 8. All piezometers were open from top to bottom, which
gives an average position of the water table in the borehole. A parallel
groundwater flow was assumed in all analyses. Table 5 shows the depth
of the piezometers and the position for low and high groundwater
level.
The data show that groundwater reacts almost immediately to rainfall inputs, suggesting that water infiltration is controlled by fissures

Table 5 Depth of piezometers and range of groundwater level position in boreholes S2,
S4, S5, S6, S9 and S11
Depth to groundwater level (m)
Borehole
Piezometer depth (m)
Minimum
Maximum

Fig. 8 Piezometric records of boreholes S2, S4, S5, S6, S9 and S11

S2
S4
S5
S6
S9
S11

8.7
7.7
8.8
19.8
9.9
7.9

0.63
0.96
3.50
3.59
1.28
0.96

6.28
4.47
6.2
11.71
5.81
5.05

Landslides 2 . 2005

89

Original Article
For this analysis, the Bishop and Janbu methods of slices have
been used by means of the STABL code, which was developed at
Purdue University, USA (Lovell 1988). The stability analyses have been
performed considering the two extreme groundwater levels recorded
in the piezometers. In this case the observed groundwater depths
ranged between 1 and 6 m in the lower unit and 10 and 12 m in the
intermediate unit. A flow parallel to the slope surface was assumed
to compute pore water pressures from water table depth. Finally,
a direct relation between water table position and global factor of
safety is defined. We have determined the factor of safety for a range
of strength parameters and water table positions. Four groups of limit
equilibrium analyses were performed:

Fig. 9 Piezometric record (blue line) and landslide velocity (black line) at borehole S2

and pipes rather than by soil porosity. The role of the karstic network in the gypsum lenses is unclear but all the observed features are
very shallow (up to 3 m depth), which is well above of the normal
groundwater level fluctuation. Thus it might play only a secondary
role.
Despite the rapid response of the piezometers, peak water levels
are attained at slightly different times depending on the permeability
of the adjacent material. Two basic types of responses to rainfall
have been observed, depending on the location of the piezometers.
These located in tension zones, such as S5, show a smaller variation
in groundwater level (ranging between 0.5 and 2 m) and a faster
drainage compared to the piezometers located elsewhere (for example
S2, S4 and S11). The latter piezometers experience changes of 2 to 5 m
and a slower rate of groundwater level decrease. The behaviour of
piezometer S5 is consistent with the presence of high permeability
cracks in the tension zone (graben). Consequently, it can be inferred
that cracks act as preferential drainage paths within the landslide.
In addition to the rapid response to precipitation inputs, all
piezometers show a defined level below which the groundwater table
decreases very slowly. This level may be observed during the periods
FebruaryApril 1997 and SeptemberOctober 1997, in which no or
negligible rain was recorded in the area (Fig. 8).
A close relationship between the groundwater level changes and
landslide activity was measured using a wire displacement meter
at borehole S2 (Fig. 9). There exists a strong level of synchronism
between the two records. In addition, the rate of displacement is
strongly correlated with the water table data. The exception to this is
the event of January 1997 (increment of velocity without increment
of water table), which may be caused by other factors (i.e. toe erosion
by the Vallcebre torrent).
Stability conditions
An analysis of the stability of the landslide was performed using 2-D
limit equilibrium method. This gives a global safety factor which is
difficult to relate to the creep behaviour seen in Vallcebre, where
movements occur almost continuously. However the determination
of both driving and resisting forces provides an insight into the magnitude of the unbalanced force and the degree of stability of the
landslide.

90

Landslides 2 . 2005

Case 1. Analyzes the stability of both the lower and intermediate


units of the landslide, i.e. the slip surfaces can include both units.
Case 2. Only the stability of the lower unit has been considered.
Case 3. Only the equilibrium of the intermediate unit has been
analyzed.
Case 4. The stability of the intermediate unit, in the absence of
the lower unit has been considered. In this case the toe of the slip
surface could exit through the crack zone, so as to avoid passive
earth pressures from the lower unit affecting the stability of the
intermediate one.
In each case, two positions of the ground water table have been
considered, corresponding to the maximum and minimum situations
measured in the piezometers within the period analyzed (Table 5).
Finally, a variation of the soil strength parameters has been adopted,
in order to perform a sensitivity analysis of the global factor of safety.
Three types of materials have been used as different layers, in
the same way as the hydrological model, with the following natural
(saturated) unit weights (n ):

Fissured shales n = 22 kN m3
Clayey siltstone with gypsum
Gravel + silty matrix


n = 20 kN m3


n = 22 kN m3

Table 6 summarises the main results of these analyses. These results


show that in some particular cases, the global factor of safety is close
to unity. This explains the activity of the landslide, i.e. small changes
in stresses or pore water pressures produce a significant decrease
in factor of safety and therefore a change in movement rate. The
geometry used in the analyses coincides with that presented in Fig. 3
(cross section A-A ). The slip surface considered in the computations
is based on the information provided by the inclinometers.
The computed factors of safety show that the lower unit of the
landslide is less stable than the intermediate one. Note that parameter
values of cohesion (c ) and friction angle ( ) used in the calculations
correspond to feasible values, taking into account the laboratory
tests performed. It has been assumed that representative strength
parameters are those of the shear surface. For the low groundwater position, displacement records at S-6 show no movement while
displacements at S-2 and S-8 are very low. This is consistent with
the results of Cases 1 and 4 which indicate that rapid movement
of the lower landslide unit removes support to the intermediate
one.

Table 6 Global Factors of Safety


computed for the different cases
(see text)

Clayey siltstone
c

Fissured
c

shale

Case 1 (Both Landslide units)


0
0
14.7
0
14.7
0
Case 2 (Lower Landslide unit)
0
14.7
0
0
0
14.7
Case 3 (Intermediate Landslide unit)
0
0
14.7
0
14.7
0
Case 4 (Intermediate unit without lower unit)
0
0
14.7
0
14.7
0

These results also agree with the field measurements, which show
higher velocities for the lower part. In fact, it is assumed that movement starts on the lower unit, which has a factor of safety close to 1,
and then a tension area and some cracks develop at the head scarp of
this unit, which is actually the toe of the medium part of the landslide. This movement of the lower unit generates, after some days, a
movement increase on the intermediate part as well. This behaviour
has been considered in the group of analyses in which the equilibrium of the intermediate part without the lower unit has been studied
(Case 4).
However, the analysis of the lower unit (Case 2) also shows that for a
low groundwater level, the factor of safety is still very low (F.S. = 0.79),
which does not correspond with the very low rate of displacement
observed. This suggests that either the analysis is not appropriate or
other forces must be taken into account.
On the other hand, Fig. 10 shows an interesting relationship between observed velocities and the depth of water table at borehole S2
for the period considered (modified from Fernandez-Pombo 1998).
A cubic curve can be fitted to the data, giving a good regression
coefficient. The points at a water table depth of 5 m and velocities
over 5108 m/s refer to the event of December 1996January 1997
mentioned above.

Colluvium
c

Factor of safety
Position of water table
Low
High

11.8
7.8

0
0

14.7
14.7

1.19
1.02

1.14
0.98

11.8
7.8

0
0

14.7
14.7

1.17
0.79

1.07
0.72

11.8
7.8

0
0

14.7
14.7

1.18
1.05

1.13
1.01

11.8
7.8

0
0

14.7
14.7

1.14
1.02

1.10
0.99

The position of the groundwater level at a depth of 6.22 m approximately, corresponds to a stable situation. If a back analysis of stability
is performed for this case in the lower unit, values of the effective
friction angle of  = 14 and of cohesion c = 0 for the fissured shale
are obtained. This, again, does not agree with the value obtained from
laboratory tests, which range from 7.8 to 11.8 . Note that this small
range of friction angle provides a change in factor of safety from 0.79
to 1.17.
The value of  = 14 obtained assuming limit equilibrium for
that position of water table is difficult to explain if only static conditions are considered. Although laboratory data could include some
uncertainties (i.e. due to sampling) we do not have any evidence of
such high strength for the slip surface. Consequently, we understand
that beside frictional resisting forces, additional resisting forces (i.e.
viscous forces) are necessary to stop the movement. Overall, a limit
equilibrium analysis can not simulate the continuous movement of
the landslide, and a different mechanical analysis must be used instead.
Modelling of landslide displacements
The close relationship existing between landslide velocity and position of groundwater level at borehole S2 suggests that it is possible
to perform a simulation of the landslide displacements from data
recorded in the piezometers. Figure 10 shows that a black box regression analysis is feasible for the simulation of displacements, but we
preferred to adopt a physical approach using the momentum equation. We analysed the dynamics of local points of the landslide in
which uniform conditions for the geometry (i.e. infinite slope conditions) can be reasonably assumed. In addition to that, a viscous
component has also been taken into account. A brief description of
the model and its application to Vallcebre is presented in this section,
which also includes a critical appraisal of the approach considered.
Viscous behaviour of the landslide
The dynamics of the landslide are governed by the difference between
destabilising forces (F), that depend basically on weight and slope,
which are constant, and resisting forces (F r ), that are sensitive to
water pressure at the slip surface. The momentum equation can be
written as:

Fig. 10 Velocities versus water table depths at borehole S2 for November 1996 to October
1998. Data correspond to mean daily values (modified from Fernandez-Pombo 1998)

F F r = ma

(1)

Landslides 2 . 2005

91

Original Article
where m is the mass and a the acceleration. For a local point of the
landslide where infinite slope conditions apply, resisting forces can
be estimated using Mohr-Coulomb criterion, depending on cohesion
and friction. Forces are computed over a unit surface, and therefore
shear stresses are considered in what follows:
[c  + ( pw ) tan  ] = ma

(2)

where is the destabilising shear stress, c is the cohesion, is the


normal stress, p w the groundwater pressure and  the friction angle,
all magnitudes referred to the slip surface. Some general predictions
on landslide behaviour can be inferred from Eq. (2) , and they will be
compared with the observations made in the Vallcebre landslide.
The groundwater pressure is the only temporal variable in the left
hand side of Eq. (2). Therefore this equation predicts a unique value
of the acceleration for each value of groundwater pressure. A one to
one relationship should also exist between acceleration and position
of the groundwater level, if parallel flow is assumed. However, the
analysis of Fig. 9 shows that the acceleration is positive when the
groundwater level rises whereas it is negative when the water table
decreases. Thus, different values of acceleration were recorded at the
landslide for the same groundwater level. This fact suggests that other
terms should be included in the momentum equation.
Another evidence of this behaviour is presented in Fig. 11. This
figure shows groundwater level changes and wire displacements at
borehole S2. Note that during dry periods, for example April and
May 1997, the movement has constant velocity. However, if there is
an unbalanced force in the system, a value of acceleration should
be expected. If this is not the case, other resisting forces should be
taken into account in Eq. (2). Therefore, an additional force that
we interpret as being a viscous component, appears to be important
and should be considered in further analyses. After a rain event,
the velocity changes, and from a total displacement BC, part can
be attributed to the rainfall (AB) and part is due to the viscous

component (AC). In fact viscous component depends on velocity,


and AC is not exactly that part, but from a conceptual point of view
it becomes evident that apart from inertial terms, other forces should
be considered in Eq. (2).
Viscous models used and simulation procedure
The mentioned evidences suggest that a viscous term should be included in Eq. (2), as it is considered in dynamics. That is:
(c  + ( pw ) tan  ) v = ma

where v is a viscous component depending on velocity.


Expression (3) has been already used to analyze landslide dynamics
by Angeli et al. (1996). They used a viscous model for a local point
analysis, assuming infinite slope conditions. A similar procedure has
been considered here. Figure 12 presents the variables used in this
approach.
The corresponding momentum Eq. (3) becomes, for infinite slope
conditions (Fig. 12):
l sin cos
[c  + ( l cos2 pw ) tan  ] v = ma

92

Landslides 2 . 2005

(4)

where is the specific weight of the sliding mass. Note that in Eq. (4),
forces are expressed per unit area of slip surface.
Viscous forces are usually dependent on the strain rate of the shear
zone. For a Bingham type model, this relation is linear and becomes:
v = v /z

(5)

where, is the viscosity, v the velocity and z the thickness of the


shear zone. Expression (5) can be introduced in Eq. (4) to give a
differential equation for a single point, assumed representative of the
whole landslide:
l sin cos [c  + (cos2 pw ) tan  ]

Fig. 11 Water table depths and displacements measured by the wire extensometer, plotted
against time at borehole S2 during spring and summer 1997

(3)

Fig. 12 Geometry and variables used in the local analyses

= ma +

dv
v
v
=m
+
z
dt
z

(6)

Pore water pressure was not measured directly, but it was estimated
from readings of depth of groundwater level. Assuming a parallel flow
to the slope surface:
pw = w cos2 h = w cos2 (l Dw )

(7)

where w is the specific weigth of water, l the thickness of the sliding


mass and D w the depth of groundwater level.
An alternative to the Bingham model based on a power law was
considered as well. In such model the velocity (v) is obtained from
the excess of driving forces 0 (Leroueil et al. 1996):

v=A

x
,

> 0

(8)

where A and x are material parameters to be calibrated by back


analysis.
In (8), 0 corresponds to the viscous component v , and that
expression can be written as:
v(power) =

 v  x1
A

(9)

By replacing in the momentum Eq. (4):


l sin cos [c  + (cos2 pw ) tan  ]
 v  x1
= ma +
A

(10)

with = l sin cos


Equations (6) and (10) have been solved numerically in terms of
displacements, using a classical finite difference scheme. Using the
same numerical approach, it is possible to derive directly the velocity
and the acceleration in the slope. After the numerical integration of
Eqs. (6) and (10), we obtained the values of the material parameters
(viscous parameters and the friction angle  ) by non linear regression
to minimize differences between the measured and the computed
displacements.
The landslide displacements used for comparison with model predictions were assessed from the displacements measured in wire extensometers. Wire displacements depend on the slope and thickness
of the basal shear zone. The slope of the shear zone was determined
from profiles including several inclinometric boreholes (see Fig. 3).
The thickness of the shear zone was obtained independently calibrating wire extensometer displacements with landslide displacements
measured with GPS. The procedure to obtain the shear zone thickness
and landslide displacements from wire extensometer displacements
is detailed in Corominas et al. (2000).
Simulated landslide displacements
An attempt to simulate the measured landslide displacements has
been performed using data recorded in several piezometers. We have
used three boreholes located in the lower unit of the landslide (boreholes S2, S9 and S11), which are considered to be representative of
different parts of the whole unit and where infinite slope assumptions

Fig. 13 Measured and predicted displacements and velocities at borehole S2 using Bingham
and power law models. The results of the two models are represented by a same line because
they are almost the same and indistinguishable at the scale of the graphs

could apply. A fourth borehole, S5, located close to the graben and
affected by some rotational component was also used. The parameters of the viscous models and the angle of friction were estimated
independently for each borehole.
Figure 13 presents a comparison between measured and predicted
displacements and velocities, using the Bingham and power law models for borehole S2. It can be observed that both models provide a
good simulation of the measured behaviour. Table 7 shows the estimated values of material parameters. The values predicted by the two
models are almost the same, because when the exponent in the power
law is near one, as occurs in this case, the two models are equivalent, and the value of /z is equal to /A. Table 7 also indicates the
goodness of fit. The latter is expressed as the root mean squared error
(RMSE).
Simulated displacements and velocities for the other two analysed
boreholes are not as good as for borehole S2. Power law model provides better fits than the Bingham model for the borehole S9, although
the difference between the velocities predicted by the two models is
small (Fig. 13 and Table 7). For borehole S11, the predictions of the
models are very similar.

Landslides 2 . 2005

93

Original Article
Table 7 Parameters obtained
from back analysis and goodness
of fitted displacements and velocities
(expressed as root mean squared
errors, RMSE's)

94

Borehole

S2

S9

S11

Thickness of shear zone (cm)


Friction angle ()
Bingham model
(kPa s)
Power law model
A (m/s)
x
Goodness of fit
Bingham model
RMSE
Power law model

31.5
7.3
1.51 107
0.55 106
1.0023
12.7
1.05 108
12.7
1.05 108

21.5
6.1
1.52 107
0.44 106
1.39
21.2
1.48 108
14.7
1.36 108

32.0
6.2
1.60 107
4.06 106
1.83
36.4
1.69 108
36.2
1.59 108

Displacement (mm)
Velocity (m/s)
Displacement (mm)
Velocity (m/s)

Fig. 14 Measured and predicted displacements and velocities at borehole S9 using Bingham
and power law models (dashed line corresponds to a period without measured data)

Fig. 15 Measured and predicted displacements and velocities at borehole S5 using Bingham
and power law models

Note that the parameters obtained through back analysis are similar
in all boreholes. On the one hand, friction angles of 6 to 7 are close
to the residual value measured in the laboratory for the fissured clay
(7.8 degrees). On the other hand, the thickness of the shear zone, a
parameter difficult to estimate in practice, is around 0.3 m, a value
obtained by analysing the behaviour of the wire extensometer records
(Corominas et al. 2000).
For the Bingham model, the viscosity calibrated was around
1.5107 kPa-s in the three boreholes. These values are within the

range of the viscosities obtained in other studies (Vulliet and Hutter 1988; Angeli et al. 1996). However, parameters estimated for the
power law provide small variations in the three boreholes, in particular the value of parameter A in borehole S11 is almost an order
of magnitude higher than the same parameter obtained in other
boreholes.
The agreement between computed and predicted displacements using the estimated parameters is generally good, particularly for borehole S2. Nevertheless at that particular point, some of the events have

Landslides 2 . 2005

not been reproduced by the model. This is the case of the peak of velocity in January 1997 without rising of water table, already mentioned.
A condition that seems to have a strong influence on the quality
of the simulation is the assumption of infinite slope. In that respect,
at borehole S2, located in the centre of the lower unit and, in a lesser
extent, at boreholes S9 and S11, these conditions may apply (Figs. 2
and 3).
Instead, for borehole S5, the infinite slope conditions are clearly
not valid. This borehole is located near the head scarp of the lower
unit where the ground surface has some rotational component of
movement. We obtained spurious results, as negative velocities or
null displacements for periods in which the landslide was moving
(Fig. 15).
The former shows that extension of the results to other local points
must be always performed carefully. The use of local models seems
to be more interesting for conceptual purposes, when the mechanical
behaviour is not known in advance and general trends are investigated as a first step for developing future global models. In this case,
the use of a viscous component in the dynamics of the landslide
has proved to be useful for a consistent analysis of the records of
displacements and water pressures. Obviously these analyses are limited to local conditions and thus, a representative set of boreholes
must be selected. For a particular landslide, those types of records
and the kind of analyses described here may become a promising
tool to develop alarm systems and to develop future procedures for
predicting landslide behaviour at a global scale (i.e., using coupled
finite element models including flow and mechanical equations).
Conclusions
The nearly continuous monitoring of the Vallcebre landslide has
allowed the observation of some characteristics of the landslide behaviour that would otherwise have gone unnoticed. The landslide is
very sensitive to rainfall, and cracks have been shown to be preferential infiltration and drainage pathways.
Our monitoring network has shown the dependence of the rate of
displacement on groundwater level; from this we infer the existence of
a viscous component in the landslide behaviour. On the other hand,
the continuous records have been used to both calibrate and validate
the analyses performed.
From the point of view of the limit equilibrium stability conditions,
the Vallcebre landslide has a factor of safety of about one. However,
when critical conditions are reached, failure does not occur instantaneously, because only a small acceleration of the mass is produced.
Thus, a full analysis needs to consider the dynamics of the landslide, instead of assuming a binary approach: stability versus failure.
The models used in the time dependent simulations are based on
this dynamic approach, rather than just considering static limit
equilibrium.
The simulations of the landslide dynamics by taking into account
a viscous component were performed using parameters obtained
by back-analysis of the recorded data. This simulation is successful
because the parameters back-analyzed are consistent with field and
laboratory data obtained in an independent manner. Moreover, those
values obtained were similar for all the boreholes considered. Thus,
the landslide behaviour at selected locations seems to be well reproduced using the procedure outlined in the paper. Next steps in this
analysis will include the examination of the behaviour of the landslide from a global perspective, instead of considering only motion
in individual boreholes. That approach, however, will likely require
a considerable computational effort, because a 3D analysis and a

coupled hydrological and mechanical model may be required. The


existence of well-instrumented landslides to implement and validate
complex numerical models is thus seen to be highly beneficial.
Acknowledgements
The financial support of this work has been provided by CEC through
the NEWTECH project (Contract ENV-CT96-0248), the Spanish CICYT (Contract AMB96-2480-CE) and the Institute of Geomodels
(IGME-DURSI-UPC-UB). The authors are indebted to David N.
Petley who has reviewed and made some valuable suggestions to this
paper
References
Angeli MG, Gasparetto P, Menotti RM, Pasuto A, Silvano S (1996) A visco-plastic model for slope
analysis applied to a mudslide in Cortina dAmpezzo, Italy. Quat J Eng Geol 29:233240
Caine N (1980) The rainfall intensity-duration control of shallow landslides and debris-flows.
Geografiska Annaler 62A:2327
Cancelli A, Nova R (1985) Landslides on soil debris cover triggered by rainstorm in Valtellina (Central
Alps, Italy). In: Proceedings of the 4th International Conference on Lanslides, Toronto, pp 262272
Corominas J, Moya J, Ledesma A, Rius J, Gili JA, Lloret A (1999) Monitoring of the Vallcebre landslide,
Eastern Pyrenees, Spain. In: Proceedings International Symposium on Slope Stability Engineering.
Matsuyama, Japan, vol. 2, pp 12391244
Corominas J, Moya J, Lloret A, Gili JA, Angeli MG, Pasuto A (2000) Measurement of landslide
displacements using a wire extensometer. Eng Geol 55:149166
Federico A, Popescu M, Fidelibus C, Interno G (2004) On the prediction of the time of occurrence
of a slope failure: a review. In: Proceedings 9th International Symposium on Landslides. Rio de
Janeiro, Taylor and Francis, London, vol. 2, pp 979983
Fernandez-Pombo MP (1998) Esllavissada de Vallcebre. Condicions dequilibri lmit del vessant i
estudi sobre el comportament viscopl`astic del moviment. Graduation Project. Civil Engineering
School, Barcelona. Unpublished
Flageollet JC, Maquaire O, Martin B, Weber D (1999) Landslides and climatic conditions in the
Barcelonnette and Vars basins (Southern French Alps, France). Geomorphology 30:6578
Fukuzono T (1990) Recent studies on time prediction of slope failure. Landslide News 4:912
Gili JA, Corominas J (1992) Aplicacion de tecnicas fotogrametricas y topograficas en la auscultacion
de algunos deslizamientos. III Simposio Nacional sobre Taludes y Laderas Inestables. La Coruna.
vol. 3, 941952
Gili JA, Corominas J, Rius J (2000) Using Global Positioning Techniques in landslide monitoring. Eng
Geol 55:167192
Griffiths DV, Lane PA (1999) Slope stability analysis by finite elements. Geotechnique 49(3):387403
Keefer DK, Wilson RC, Mark RK, Brabb EE, Brown WM, Ellen SD, Harp EL, Wieczorek GF, Alger CS,
Zatkin RS (1987) Real-time landslide warning during heavy rainfall. Science 238:921 925
Larsen MC, Simon A (1993) A rainfall intensity-duration threshold for landslides in humid-tropical
environment, Puerto Rico. Geografiska Annaler 75A(12):1323
Lateltin O, Bonnard Ch (1995) Reactivation of the Falli-Holli landslide in the Prealps of Freiburg,
Switzerland. Landslide News 9:1821
Leroueil S, Locat J, Vaunat J, Picarelli L, Lee H, Faure R (1996) Geotechnical characterization of
landlsides. In: Senneset K (ed) 7th International Symposium on Landslides. Trondheim, A.A. vol.
1. Balkema, Rotterdam, pp 5374
Lovell CW (1988) Users guide for PC STABL 5M. Purdue University, Indiana
Matsui T, San KC (1992) Finite element slope stability analysis by shear strength reduction technique.
Soils Found 32:5970
Noverraz F, Bonnard Ch, Dupraz H, Huguenin L (1998) Grands glissements et climat. Rapport final
PNR 31. Vdf. Zurich. 314 pp
Petley DN, Allison RJ (1997) The mechanics of deep-seated landslides. Earth Surf Proc Landforms
22:747758
Petley DN, Bulmer MH, Murphy W (2002) Patterns of movement in rotational and translational
landslides. Geology 30(8):719722
Picarelli L, Urciuoli S, Russo C (2000) Mechanics of slope deformation and failure in stiff clays and clay
shales as a consequence of pore pressure fluctuation. Proceedings 8th International Symposium
on Landslides, Cardiff, vol. 4, 34 pp
Saito M (1965) Forecasting the time of occurrence of a slope failure. Proceedings 6th International
Conference on Soil Mechanics and Foundation Engineering. Montreal, vol. 2, 315318
Van Asch ThWJ, Buma JT (1997) Modelling groundwater fluctuations and the frequency of movement
ofalandslideintheTerresNoiresofBarcelonnette(France).EarthSurfProcLandforms22:131141
Voight B (1989) Material science law applies to time forecast of slope failure. Landslide News 3:811
Voight B, Kennedy J (1979) Slope failure of 1967-1969 Chuquicamata mine, Chile. In: Voight B (ed).
Rockslides and avalanches. Developments in geotechnical engineering 14B. Elsevier, pp 595632
Vulliet L (2000) Natural slopes in slow movement. In: Zaman M, Gioda G, Booker J (eds). Modeling
in geomechanics. John Wiley, Chichester, pp 654676

Landslides 2 . 2005

95

Original Article
Vulliet L, Hutter K (1988) Some constitutive laws for creeping soil and for rate-dependent sliding at
interfaces. In: Swoboda (ed) Numerical methods in geomechanics (Innsbruck 1988). Balkema,
Rotterdam, pp 495502
Wieczorek GF (1987) Effect of rainfall intensity and duration on debris flows in central Santa Cruz
Mountains, California. Geological Society of America. Rev Eng Geol 7:93104
Wilkinson PL, Anderson MG, Lloyd DM (2002) An integrated hydrological model for rain-induced
landslide prediction. Earth Surf Proc Landforms 27:12851297

96

Landslides 2 . 2005

J. Corominas () J. Moya A. Ledesma A. Lloret J. A. Gili


Department of Geotechnical Engineering and Geosciences, Technical University of Catalonia, UPC,
Jordi Girona 1-3, D-2 Building,
E-08034 Barcelona, Spain
e-mail: Jordi.Corominas@upc.edu
Tel.: +34-93-401-6861
Fax: +34-93-401-7251

You might also like