0% found this document useful (0 votes)
126 views

Best Practice PDF

Uploaded by

michel
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
126 views

Best Practice PDF

Uploaded by

michel
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 61

GEOLOGICAL SURVEY OF CANADA

OPEN FILE 8114

SITE INVESTIGATION, ANALYSIS, MONITORING AND


TREATMENT
Canadian Technical Guidelines and Best Practices related
to Landslides: a national initiative for loss reduction

M. Lato, P. Bobrowsky, N. Roberts, S. Bean, S. Powell, S. McDougall,


M-A. Brideau, D. Stead and D. VanDine

2016

GEOLOGICAL SURVEY OF CANADA


OPEN FILE 8114

SITE INVESTIGATION, ANALYSIS, MONITORING AND


TREATMENT
Canadian Technical Guidelines and Best Practices related to
Landslides: a national initiative for loss reduction
M. Lato1, P. Bobrowsky2, N. Roberts3, S. Bean4, S. Powell4, S. McDougall5,
M-A. Brideau3, D. Stead3 and D. VanDine6
1

BGC Engineering Inc., Ottawa, Ontario


Geological Survey of Canada, Sidney, British Columbia
3
Simon Fraser University, Burnaby, British Columbia
4
Thurber Engineering Ltd., Victoria, British Columbia
5
University of British Columbia, Vancouver, British Columbia
6
VanDine Geological Engineering Ltd., Victoria, British Columbia
2

2016
Her Majesty the Queen in Right of Canada, as represented by the Minister of Natural Resources, 2016
Information contained in this publication or product may be reproduced, in part or in whole, and by any means,
for personal or public non-commercial purposes, without charge or further permission, unless otherwise
specified.
You are asked to:

exercise due diligence in ensuring the accuracy of the materials reproduced;

indicate the complete title of the materials reproduced, and the name of the author organization; and

indicate that the reproduction is a copy of an official work that is published by Natural Resources
Canada (NRCan) and that the reproduction has not been produced in affiliation with, or with the endorsement
of, NRCan.
Commercial reproduction and distribution is prohibited except with written permission from NRCan. For more
information, contact NRCan at [email protected].
doi:10.4095/299117
This publication is available for free download through GEOSCAN (http://geoscan.nrcan.gc.ca/).
Recommended citation
Lato, M. et al. 2016. Site Investigation, analysis, monitoring and treatment Canadian technical guidelines and
best practices related to landslides: a national initiative for loss reduction; Geological Survey of Canada,
Open File 8114, 59 p. doi: 10.4095/299117
Publications in this series have not been edited; they are released as submitted by the authors.

ii

Canadian Technical Guidelines and Best Practices related to Landslides: a


national initiative for loss reduction
SITE INVESTIGATION, ANALYSIS, MONITORING AND TREATMENT
Note to Reader
This is the final publication in a series of Geological Survey of Canada Open Files that have
appeared during the past several years. The series forms the basis of the Canadian Technical
Guidelines and Best Practices related to Landslides: a national initiative for loss reduction.
Once all Open Files have been published, they will be further edited, compiled and published
as a GSC Bulletin. Each open file in the series corresponds to a chapter in the bulletin.
Comments on this Open File, or others in this series should be sent to Dr. P.
Bobrowsky at [email protected].

1.0 INTRODUCTION
The purpose of this contribution is to present a what to do and not how to do discussion
regarding the investigation, analysis, monitoring and treatment of landslides. It presents, in
general terms, what to consider when planning and carrying out site investigations, when
selecting appropriate methods to analyze slopes for stability, when designing and
implementing a monitoring program, and when planning and designing the treatment of an
existing or potential landslide (see also Highland and Bobrowsky, 2008). The Open File
reviews the fundamentals of such topics, but only provides broad details of methods and
techniques, their appropriate use and limitations in summary tables with appropriate
references. The methods and techniques associated with all the topics in this contribution are
developing rapidly, and landslide professionals should keep themselves current through
continuing professional development. It is not the intent of this chapter to reproduce the
details of the methods and techniques that make up the topics of investigation, analysis,
monitoring and treatment. Terminology follows Bobrowsky and Couture (2014) for all
contributions in this series.

2.0 PRE-PLANNING
The following sections will more fully develop site investigation, analysis, monitoring and
treatment related to landslides and potential landslides, however, some initial work is
routinely required before a site investigation is launched including the subsequent tasks of
analysis, monitoring and/or treatment begin.
As discussed in VanDine (2012), the initial step of the risk management process follows
the recognition that a landslide has occurred, or could occur, and that human health and
safety, aspects of the environment and/or financial interests have been, or could be, affected.
During this step, possible landslide risk scenarios and stakeholders should be identified.
(Stakeholders are persons and organizations, including government agencies that can affect,
be affected by, or perceive themselves to be affected by the landslide, or by associated
1

decisions or activities.) In addition, early in the process the scope, goals and potential methods
of managing the landslide risks should be established by the landslide professional.
Elsewhere VanDine (2011) reviews the aspects that landslide professionals should
minimally undertake prior to a landslide study.
In an emergency response situation, some of the prerequisites proposed by VanDine are
not practical or cannot be achieved in a timely manner. In such situations, a landslide
professional and their client should at least have an agreement as to the objectives and the
deliverables of the landslide study.
The level of effort of a landslide study can range from overview to detailed, and should
be determined, relative to the project objectives and intended use of the results, study area,
complexity of the terrain, elements at risk, and available background information. Because of
the unknowns involved, it is often useful to use a phased study approach trending from
overview to more detailed.
The study area should be determined by the type of landslide, the geological and
geotechnical complexity of the terrain involved, and the elements at risk. It should not be
limited to a lot, property or political boundary, but should include other areas that could
potentially affect, or be affected by, landslide. Some types of landslides can travel long
distances, and therefore, where appropriate, the study area should include sources areas, travel
paths and depositional areas.

3.0
3.1.

SITE INVESTIGATIONS
Background

Prior to any site investigation, the landslide professional should collect, possibly with the help
of stakeholders, available existing information associated with the study area. During the
initial phase of work landslide risk scenarios should be identified alongside the time frame in
which the work should be conducted. The landslide professional should consider the
following items and their respective levels of reliability, as possible sources of existing
information (cf. Jackson et al., 2012):
large and small scale topographic and cadastral maps;
airphotos of different years (historical to present) and scales;
bedrock, geomorphology and surficial geology;
terrain maps, terrain stability maps, landslide inventories, landslide hazard maps
and reports;
seismic data, where appropriate, including: seismic hazard maps and reports;
ground motion data, seismic Site Class, and modal magnitude values of the design
earthquake;
water well, borehole and other subsurface drill records and hydrogeology reports;
flood plain mapping and alluvial fan mapping;
maps that show existing and proposed development, infrastructure such as
transportation routes, utilities, surface drainage, in-ground disposal of storm water,
and in-ground disposal of waste water and/or sewage;
2

in areas of natural resource activity, such as forestry, mining, oil and gas, and
hydro development: resource inventory maps and reports, resource development
plans, watershed and terrain stability assessments, past and proposed resource road
construction and other activities, and other relevant resource activity information;
urban, planning and other anthropogenic activity reports;
evidence and history of wildfires and other changed conditions in the area, and;
previous geological, geotechnical and landslide reports that address the study area
and, if available, neighbouring areas.
Information can be obtained from published and non-published sources through various
federal and provincial agencies, local governments and other sources. For larger areas,
obtaining project-specific information, in addition to existing background information may be
useful. Examples include airphotos, high-resolution satellite imagery, and LiDAR (Light
Detection and Ranging) images that may be used for geological and geomorphological
mapping and/or topographical mapping.
Background information should be reviewed prior to undertaking subsequent phases of
the site investigation, analysis, monitoring and treatment. If information is known to be
available and the landslide professional did not or was not able to obtain it, this fact should be
noted (Gerath et al., 2010).
3.2.

Data Information Management

Organized data management the collection, storage, analysis and presentation of data is
important in geosciences due to the extensive use of geospatial data. Landslide investigations
typically necessitate data structures that are three-dimensional. Some investigation types, such
as monitoring or multi-temporal inventories, also require data to be structured temporally.
With the advent and development of the field of geomatics, data management has become
increasingly important.
New and emerging technologies, particularly remote sensing techniques, expand
opportunities to characterize landslides and increase the accuracy and precision of individual
datasets (cf. Petley, 2012; van Westen et al., 2008; Delacourt et al., 2007). Consequently, data
volume and computational need also increase. Data management has thus changed
dramatically over recent decades, with growing emphasis on geographic information systems
(GIS). This section summarizes relevant low-cost datasets available for Canada and provides
guidance to ease the combination, analysis and presentation of datasets typical in landslide
investigations though planning of data management.
3.2.1. Compile and collect
Various considerations are necessary during capture of data to facilitate their subsequent
integration into geospatial databases. An overview of important considerations for
compilation of existing datasets, and collection of new datasets is given here. Figure 1
provides access sources and general details for basic datasets, of regional or greater extent,
that are useful in landslide investigations in Canada.
Several options exist for most data types; datasets that are selected for compilation and
collection will be a function of budget, temporal and spatial data availability, and
investigation objectives. Identification and, when necessary, compilation of previously
3

collected datasets should be undertaken early in the investigation. Knowledge and use of these
reduces time and resources needs by preventing duplication of datasets as well as by
providing a basis for preliminary analysis and guidance for further data collection.
3.2.1.1. Information datasets
Topography and geology datasets are derived from pre-existing sources. For many parts of
Canada they are now available in digital georeferenced forms from federal and provincial
sources.
3.2.1.1.1 Topographic datasets
Pre-existing topographic datasets are produced from remotely sensed datasets and, therefore,
suffer from issues inherent in the imagery from which they were derived. Government-issued
topographic datasets are widely available at small scales (~1:125,000) to large scales
(~1:5,000), although larger scale datasets are typically not free. All are arranged by tiles and
are available for download. Provincial and territorial datasets typically have greater detail
compared to regionally extensive topographic datasets.
Canadian Digital Elevation Database (CDED) DEMs are compiled from hypsographic
and hydrographic elements of the NTDB (Centre for Topographic Information, 2000), which
are in turn based on stereo aerial photographs. These are available as 1:50,000 and 1:250,000
NTS tiles. Contour lines are included in the NTDB, but they are often misaligned between
tiles given different contour intervals or, in the case of some older map sheets, use of imperial
elevation units.
Most DEMs are provided in geographic coordinate systems. They must, therefore, be
converted to a metric-based projection before slope angle and hill shade tools in GIS packages
can be used to produce properly scaled results.
Digital elevation models express surfaces as regularly spaced grids. Consequently, they
are poor at representing abrupt topographic changes that are muted due to the averaging effect
between cells. The use of a higher spatial resolution DEM, even of lower accuracy, can
reduce this effect (Environment Yukon, 2012). Alternatively, the surface model can be
converted to a triangular irregular network (TIN) (Peucker et al., 1978) with use of break
lines.
3.2.1.2. Geologic datasets
Digital bedrock geology is available for most regions from provincial or territorial
governments. Specialized reports produced by Natural Resources Canada (NRCan) and
provincial ministries, as well as by some consultants, provide additional details on the
geology of select areas.
Where local-scale maps exist in non-geospatial forms, they can be scanned and
georeferenced for use as images. Alternatively, they can be digitized for use as vector data,
which increases their utility in analysis; although labour intensive. Some Geological Survey
of Canada (GSC) Open File reports provide digitized versions of past studies (e.g. Dunn and
Ricketts, 1994).

3.2.2. Remote datasets


3.2.2.1. Aerial photography
Despite the advent of sub-metre-resolution optical satellite sensors during the past decade,
conventional aerial photography remains the highest image resolution option and is an
excellent source of historic information. The majority of aerial photographs available for
Canada are from federal and provincial acquisitions. Typical scales are 1:15,000 (low
altitude), 1:40,000 (moderate-altitude) and 1:60,000 (high-altitude). Aerial photograph
availability is greatest for developed areas, where the earliest photography dates from the
1920s and with frequent subsequent acquisition. Surveys for most non-developed areas range
in frequency from half a decade to several decades. Additionally, special surveys typically
providing low-altitude flights for specific areas of interest (e.g. resource management or
engineering projects) are sometimes available.
Some Canadian aerial photograph sources are transitioning to digital query and delivery.
Federal aerial photographs can be acquired through the National Air Photo Library (NAPL).
Provincial ministries are typically the best source for provincial aerial photographs.
Acquisition of new federal or provincial aerial photography is limited.
Spatial distortions in aerial photographs are of two types: radial displacement, which
increases systematically away from the nadir location, and terrain displacement, which is
irregular. The distortions result in a variable scale over a single photograph that prevents its
spatially accurate inclusion in a GIS. Orthophotographs, which produce orthographic view
though removal of radial and terrain displacement effects, are available for many regions of
Canada. They are suitable for display and analysis in a GIS, but lack stereo capability, which
is of great importance for landslide investigations.
Two options exist for transfer interpretations from aerial photographs into a GIS.
Interpretations made from hard-copy photographs can be subsequently digitized in a GIS,
using a georectified reference based layer (e.g. orhthophotography, digital satellite image, or
large-scale topographic map). Alternatively, specialized software packages now allow
simultaneous soft-copy stereo-interpretation and direct digitizing of features of interest into a
GIS. These require ground control points and details of the camera used.
3.2.2.2. Space-borne optical imagery
Since the 1970s a wide variety of space-borne optical imaging systems have been developed,
with image resolution and temporal coverage increasing over time. The Landsat satellite
series provides the longest record starting in 1972. Landsat imagery is useful for moderateresolution needs, particularly where multispectral processing is required. Freely available
orthorectified scenes for all parts of the globe are available for download.
Higher spatial-resolution satellites typically offer only a few multispectral bands.
Relative to Landsat-7, ASTER offers improved (15-m) resolution in three spectral bands.
Scenes are available since 2000 for a nominal cost. The Systme Pour lObservation de la
Terre (SPOT) satellite series provides moderate-resolution to high-resolution imagery from
1986 onward; panchromatic resolutions ranging from 2.5 m to 10 m and multispectral
resolutions ranging from 10 m to 20 m. SPOT 4 and SPOT 5 are currently operational, SPOT
6 (launched 9 September 2012) provides 1.5 m panchromatic and 8 m multispectral imagery.
5

The RapidEye constellation consists of five satellites offering 5 m resolution multispectral


imagery with a 24-hour revisit period, designed for rapid disaster response (RapidEye, 2012).
A growing number of very high-resolution commercial satellites offer high-cost sub-metre
imagery. Current systems include Ikonos, GeoEye-1, Quickbird, Worldview-1 and
Worldview-2.
Prior to ordering, the level of processing and type of optical satellite data must be
selected. Three levels or processing are typically available: images corrected only for
geometric and radiometric distortions of the sensor; georectified images; and orthorectivited
images. To reduce cost, it is possible to purchase pan-sharpened multispectral imagery, which
combines the spatial resolution of a panchromatic band with multispectral data (Lillesand et
al., 2008). However, doing so greatly limits multispectral processing options.
Imagery can be queried and ordered form the satellite operators website as well as
numerous contractors websites. Many full scene images can also be viewed in Google Earth,
allowing free and immediate interpretation of many landslide features, but without the
benefits of image analysis techniques.
Most high-resolution and very high-resolution satellites can also be tasked to acquire
imagery of specific areas within specific date windows, offering high flexibility in data
capture.
3.2.2.3. RADAR
Orbital RADAR sensors are fundamentally different from optical sensors. They are not
limited by time of day or cloud cover due to their active nature and their use of microwaves,
respectively. RADAR images represent surface texture, structural features and ground
moisture. Thus, RADAR is suitable for detection of recent landslides, landslide-dammed
lakes, as well as geologic and hydrologic conditions that may be related to landslide
occurrence.
Additional applications relevant to landslide investigation are provided by RADAR
interferometry (InSAR), which determines the non-random phase difference between
RADAR scenes and displays the differences as an interferogram (Massonnet and Feigl, 1998).
Site-specific DEMs can be generated with repeat-pass InSAR (Massonnet and Feigl, 1998).
Differential InSAR, which uses either a synthetically generated DEM or, more often,
additional RADAR scenes to isolate phase differences due to ground motion (Massonnet and
Feigle, 1998), is commonly used in displacement detection of slow-moving landslides.
Factors affecting RADAR imagery are especially problematic in mountainous terrain.
Foreshortening causes geometric compression of data whereas layover and shadow cause
complete loss of data (Lillesand et al., 2008). Severity of these factors increases with terrain
steepness, requiring incidence angles to be carefully chosen. Artifacts problematic in InSAR
can result from vertical atmosphere stratification in areas of high topographic variability
(Hanssen, 2001). Most RADAR sensors offer a range of beam modes with a trade-off
between resolution and scene footprint.
3.2.2.4. LiDAR
Data from LiDAR surveys are useful in many aspects of detailed landslide site investigations,
including generation of DEMs or TINS as well as mapping of geomorphic and structural
6

features. As data collection increases, repositories of LiDAR in Canada will expand, similar
to those already available in the United States (e.g. Puget Sound LiDAR Consortium,
National Oceanic and Atmospheric Administrations [NOAA] coastal LiDAR collection).
Appropriate filtering is necessary to remove non-ground laser pulse returns from 3D
point clouds. The resulting bare-earth LiDAR surfaces, represented by the last laser pulse
returns, are typically of greatest use in landslide investigations. However, in areas of very
dense vegetation, laser pulse may not reach the ground surface, resulting in vegetation-related
artifacts in the bare-earth surface (e.g. Zhang et al., 2003).
3.2.2.5. Field methods
The various types of field data generated during an investigation may be integrated by manual
or automated means with remotely sensed and archival geospatial data.
In the case of field traverses or geophysical surveys, GPS tracking of data collection is
advisable. Data and interpretations can be coded to appropriate locations using attribute tables
in a GIS, whereas photographs or other images can be hyperlinked. Data recorded by
permanent or semi-permanent instrumentation can be recorded and transferred intermittently
to a GIS or, if communications options are available, transferred in real time. Products of
terrestrial remote sensing techniques can be merged into a GIS by converting data into an
appropriate map projection.
3.2.3. Store and analyze
3.2.3.1. Geospatial databases and Geographic Information Systems
Geospatial databases are structured sets of data with referenced locations on (or near) the
Earths surface. They are optimized for organization and integration of datasets from different
sources, as well as storage and structuring of datasets in a standardized way to ease sharing by
users. Geographic Information Systems simply combine a geospatial database with tools to
conduct analyses and produce new data and information. Consequently, GIS has become the
standard tool for data management in the geosciences.
Regardless of the software package or packages used, two simple steps will avoid some
problems commonly encountered with GIS. Use of the same projection for all datasets will
mitigate misalignment issues. Doing so will likely require conversion or even definition of
projections for some datasets. Once aligned, all datasets can easily be re-displayed within a
GIS using other projections. Selection of a projection depends on spatial scale, location, and
shape of the area considered. Site-scale projects commonly use the appropriate grid zone of
Universal Transverse Mercator projection, whereas national-scale projects commonly use
Lambert conformal conic. Projections used for regional-scale and provincial-scale projects are
varied.

Figure 1. Geospatial datasets with respect to various scales of data available and their typical
use.
3.3.

Site investigation techniques

Subsurface intrusive sampling and testing techniques help geoscientists and engineers assess
the nature and extent of stratigraphy and mechanical properties of the various units of soil and
rock present below surface (Jackson et al., 2012; Johnson and DeGraff, 1988). These, along
with surface mapping and geophysical studies, are the critical building blocks for a good
geological model to be used in subsequent slope stability analyses and the design of
remediation structures. Supplementary information about the sampling and testing techniques
discussed here (see Table 1) can be found in most landslide investigation and engineering
geology textbooks (e.g. McGuffey et al., 1996; Wyllie and Mah, 2004; Cornforth, 2005; Hunt,
2007; Gonzalez de Vallego and Ferrer, 2011). Since subsurface intrusive methods sample a
8

small volume of soil or rock, they are often used in conjunction with surface geophysical
techniques to extend their applicable range (Bichler et al., 2004).
Table 1. Example of site investigation techniques.
Method
Surface
Aerial extent of landslide
Characterize elements at risk
Structural characterization
Soil sampling
Shear surfaces
Subsurface
Borehole logging
Borehole televiewer
Packer test
Geophysical methods
Trenching
Penetration testing
Shear vein testing

Application

Define the spatial extents, area and volume of the landslide


region
Identify and categorize the elements at risk and their spatial
relationship with the landslide
Map the structural condition of the landslide to understand the
failure mechanism and movement history
Identify different material present and potential for different
failure mechanisms
Identify failure surfaces and movement history

Collect intact soil and rock samples


Image walls of boreholes to map discontinuity orientations and
conditions
Measure water quality and hydraulic conductivity
Measure subsurface conditions, features, and material types
Map material type, thickness and lateral extent. Retrieve samples
for testing
Map stratigraphy and mechanical properties of material terrestrial
and submarine environments
Measure the torsional force required to shear a cylindrical volume
of soil

Groundwater
Ground water level
Slope of ground water table
Hydraulic gradient
Pore pressure
Groundwater flow
Lab testing
Unconfined Compressive Strength Measure the uniaxial strength of a rock or soil material
Direct shear test
Measure the shear strength of a rock or soil material
Liquid limit
Measure the water content when the behaviour of a clayey soil
changes to liquid from plastic.
Water density

3.3.1. Sampling
3.3.1.1. Boreholes
Borehole investigations cover a wide range of different techniques to excavate a small
(typically a few centimetres) diameter cylindrical hole in soil or rock masses. The specific
9

method used depends on the expected material type, need for intact sample, depth of
investigation, and topography at the site of interest. Auger drilling is a low cost, portable and
rapid method for soil investigation that consists of driving a helical screw into the ground by
hand or motorized equipment. BGC Engineering Inc. (2005) used a hand auger to assess the
location and depth of each layer of surficial material for a landslide risk assessment along the
Berkeley escarpment, Vancouver, B.C. Diamond drilling is a popular technique used to
investigate soil and rock masses as is it collects continuous samples and typically has high
recovery rate. A motor powers a rotating diamond tip bit which is attached to hollow drilling
rods that collect undisturbed samples. Details about this and other drilling techniques and
geotechnical core logging can be found in various textbooks (e.g. Wyllie and Mah, 2004;
Hunt, 2007).
Downhole optical and acoustic televiewers are tools that image the wall of a borehole and
allow information about discontinuity orientations to be acquired. This approach was used at
Turtle Mountain, Alberta by Spratt and Lamb (2005) to obtain discontinuity orientations and
image dissolution cavities. Various downhole geophysical techniques (natural gamma,
magnetic susceptibility, electrical conductivity, temperature, and seismic) can be used to
better define the physical properties of the various lithological units identified (for more
details see Sassa, 2006 and Dunnicliffe, 1993). Borehole geophysical techniques were used by
Aylsworth and Hunter (2003) in the Ottawa Valley Landslide Project to provide information
about critical parameters related to regional slope stability conditions.
3.3.1.2. Trenches
Trenches are mechanical (e.g. back hoe) or manual (e.g. spade) trial excavations. They are
quick and inexpensive methods to investigate surficial material allowing for the composition,
thickness and lateral extent (limited, but better than for boreholes) to be described and the
material to be tested and sampled. For a safe site investigation, care must be taken for proper
support of the trench wall. Prior to excavation, provincial or territorial occupational health and
safety (OHS) regulations and standards should be consulted and followed. Trenches have
been excavated across antislope scarps at Mount Curie and Handcarp Peak Curie to
investigate the stratigraphy of the sediment filling the features and infer the history of
movement (Thompson et al., 1997; Hensold, 2011).
3.3.2. Testing
3.3.2.1. Penetration tests
Penetration tests measure the soil resistance to the drive of a standard size sampler or probe.
The two most popular of these tests are the standard penetration test (SPT) and the cone
penetrometer test (CPT). The SPT test consists of counting the number of blows per 15 cm
advance of a split-spoon sampler whereas the CPT test measures the tip resistance and sleeve
friction of a probe as it is pushed continuously into the ground. The penetration tests (SPT and
CPT) have been used to provide information about stratigraphy and mechanical properties of
the material in the subsurface at various terrestrial (e.g. Mahmoud et al., 2000) and submarine
(e.g. Mosher et al., 2004) landslides in Canada. Standard procedures have been proposed for

10

these in-situ penetration tests (e.g. ASTM, 2005; 2007; 2008a; 2011). The guideline that is
followed should be specified when reporting test results.
3.3.2.2. Shear vanes
The shear vane test consists of inserting a four-bladed vane into an undisturbed clay-rich soil
horizon and measuring the torsional force required to shear a cylindrical volume of soil. The
maximum torsional force applied during the test can be related to the undrained shear strength
of the soil. The testing procedure and correlations are presented in ASTM (2008b). Shear
vane tests have been used to assess the strength of sensitive glaciomarine clays from the
Ottawa-region (Crawford and Eden, 1967) and from central British Columbia (Geertsema and
Torrance, 2005).

4.0

SLOPE ANALYSIS

4.1.

Introduction

A slope analysis is usually carried out after a site investigation has been completed. It can be
carried out either as a result of a landslide occurring, or before a potential landslide occurs. If
carried out after a landslide has occurred, a slope analysis typically tries to determine the type
of movement, geological conditions, geotechnical properties prior to the landslide and/or the
preparatory causes and triggering factors, and/or to estimate the current stability of the slope
to assist in the evaluation and design of treatment, either mitigation or remediation (see
Section 6).
A slope analysis of a potential landslide is typically used to estimate how likely it is that a
landslide will occur and/or where it will travel or retrogress to if it does occur; in other words,
a hazard analysis. There are a number of methods to estimate the likelihood of occurrence that
includes risk analysis, limit equilibrium analyses, slope displacement analysis and slope
deformation analyses. Similarly there are a number of methods to estimate where a landslide
will travel that include empirical analysis, analytical methods, physical models and numerical
methods.
This section briefly describes a number of methods to estimate how likely it is that a
landslide will occur and where it will travel or retrogress to if it does occur. The methods used
depend on i) the types of movement involved or potentially involved and ii) the types of
material involved.
The method of analysis and the level of effort should also be consistent with the purpose,
scope and other requirements of the project and should be appropriate to size of the study
area. Typically the greater the perceived hazard and risk, the greater the level of effort
expended. Other factors include the relative stability; elements at risk; geological and
geotechnical complexity of the terrain; availability, quality and reliability of background
information and field data; and the tolerable and/or acceptable risk criteria that are available
and/or are appropriate.

11

4.2.

Methods

4.2.1. Risk Analysis


Risk analyses can be carried out for all types of landslides. In its simplest form risk (R) is the
product of probability or likelihood of landslide occurrence (P) and consequence (C).
R=PxC

Equation 4-1

In a more complex, but more realistic and practical form, risk is the product of four
components as shown in Table 2.
Table 2. Four components of risk (adapted from BC MOF, 2002).
Component
P(H), probability
of occurrence
P(S:H), spatial
probability
P(T:S), temporal
probability

V(T:L),
vulnerability

Description
-probability or likelihood of a landslide occurring
-will the landslide affect a specific location?; considers where a landslide will travel
or regress to
-if an element at risk is mobile, will it be at that specific location at the time of the
landslide?
-if an element is not mobile, it will be at that specific location at all times, therefore
P(T:S) = 1
-if the element at risk is at that specific location at the time of the landslide, its
vulnerability depends on the type and the character of the element; its robustness (or
fragility) and its exposure to (protection from) the landslide;
-if total loss of an element is assumed, V(L:T) = 1

Putting these four components together,


R = P(H) x P(S:H) x P(T:S) x V(L:T)

Equation 4-2

If P(T:S) and V(L:T) are assumed to be 1 (refer to Table 2, above),


R = P(H) x P(S:H)

Equation 4-3

Such assumptions must be made only after due consideration of the consequences, and
when the ramifications on the specific project are considered and discussed with the
stakeholders. The concept of risk analysis is developed further in Wise et al. (2004, Chapter
3), Fell et al. (2005), AGS (2007) and Porter and Morgenstern (2013).
Common methods of estimating probability (and risk) include, risk matrices (risk bins),
event tree decomposition, and quantitative risk analysis (QRA). Each method has its
advantages and disadvantages. These, along with references to some examples, are presented
in Table 3.

12

Table 3. Some Methods of Risk Analysis, advantages and disadvantages and examples with
references.
Method
Risk Matrices
(Risk Bins)

Advantages/Disadvantages
-relatively simple method of combining two
components of risk; typically associated with
qualitative risk analysis

Example
Wise et al. (2004, Chapter 3)

Event Tree

-relatively simple and thorough method of


decomposing the risk components; typically
associated with quantitative risk analysis, but
can provide a helpful framework for qualitative
risk analysis, particularly for more complex
projects

Wise et al. (2004, Chapter 3)

Quantitative
Risk Analysis
(QRA)

-relatively complex method that can be used to


incorporate just some or all of the components
of risk; as the name implies, this is always
associated with quantitative risk analysis; can be
used to develop F-M (frequency-magnitude) or
F-N (frequency-number of fatality plots)

Friele and Clague, 2005;


BGC Engineering Inc (2006)

From the beginning, the appropriate method of risk analysis must be selected and used.
The method should be appropriate to the:
situation and purpose of the analysis
level of risk and the elements at risk or potentially at risk
availability, quality and reliability of data, and
form of the acceptable or tolerable threshold of risk.
Quantitative methods of risk analysis use numerical values, or ranges of numerical
values, to express the various components of risk. Qualitative methods use terms, such as very
high, high, moderate, low and very low, to express relative values or ranges of values. There
are no standard definitions for terms that express qualitative relative values and to avoid
ambiguity the terms should be defined. Some examples of qualitative terms are provided in
Table 4. Other examples are presented in Appendix 10 of BC MOF (2002), Tables 2 to 7 of
Chapter 3 and Appendix 5 of Wise et al. (2004) and Appendix B of AGS (2007).
Table 4. Examples of qualitative terms and quantitative values.
Example
Annual probability of a landslide
occurring

Qualitative Term
Low

Quantitative Value
1:500 to 1:2500 (0.002 to 0.0004)

Probability of a landslide occurring


along a forest road

Moderate

18% to 64% over the design life of


the road (0.18 to 0.64)

Probability of a landslide reaching


a specific location

High

75% (0.75)

With regard to accuracy, quantitative estimations may be no more accurate than


qualitative estimations. The accuracy does not depend on the use of numbers, but rather it
13

depends on whether all the components of the analysis have been appropriately considered
and the availability, quality and reliability of the required data. The term semi-quantitative is
sometimes used to describe a combination of quantitative and qualitative analysis methods.
This term is inappropriate for use in risk studies, and when quantitative and qualitative
estimations are combined, the results are more appropriately referred to as qualitative.
For natural processes that occur frequently in the same location, such as floods, snow
avalanches, and some debris flows, probabilities of occurrence can be estimated by rigorous
statistical analysis. For example, after measuring the annual flood stage on a river for many
years, the flood stage of the 200-year flood on that river can be estimated objectively.
Objective probability can also be used to estimate spatial and temporal probabilities, and
vulnerability. Objective probability estimates assume that past events, and their pre-cursor
conditions, are reasonable predictors of future conditions. This in itself is a subjective
assumption, and therefore, no estimate is entirely objective.
Most landslides, however, rarely occur frequently at the same location and therefore it is
difficult to objectively estimate probability of occurrence. Probabilities of occurrence of
landslides are often estimated subjectively. Subjective probability is a measure of one's belief
that a landslide will occur. It is based on empirical evidence combined with professional
judgment. Subjective probability estimates are no less valid than objective probability
estimates, provided the scientific basis for the former are well explained. Subjective
probability can also be used to estimate spatial and temporal probabilities, and vulnerability.
Generally, the more knowledgeable and experienced the landslide professional, the more
reliable are their subjective professional judgments. With respect to the probability of
occurrence, for example, knowledge and experience include:
landslide-specific surface and subsurface observations, field and laboratory testing
results and instrumentation/monitoring results
reliability and applicability of landslide models
results of any other methods of slope analysis (e.g. factor of safety, slope
deformation, modelling),
published case histories, and
general and local personal experience.
Objective probabilities are usually expressed quantitatively. Subjective probabilities can
be expressed either quantitatively or qualitatively. As with objective probability, subjective
probability should consider a specified time period, for example, a year or the design life of
the element at risk, and the likely site and weather conditions over that period. Because no
landslide professional has complete knowledge or perfect judgment, all subjective
probabilities contain some uncertainty, and typically lie between 1 and 0.
Vick (2002) is a good reference on the topic of subjective probability and engineering
judgment, and includes some techniques of estimating subjective probabilities and some tips
for assigning both quantitative and qualitative subjective probabilities for landslide risk
analyses. These are also summarized in Wise et al. (2004; Chapter 3).
4.2.2. Limit Equilibrium Analysis
Limit equilibrium analysis involves estimating a factor of safety (FS) for a specific rupture
surface. A FS is the ratio of forces on a slope that resist movement (resisting forces) to those
that promote downslope movement (driving forces). Examples of resisting forces include
14

friction and cohesion. Examples of driving forces include static forces, for example gravity,
and dynamic forces, for example earthquake loading.
FS = resisting forces/driving forces

Equation 4-4

Limit equilibrium methods tend to be most suited to slide modes of movement that
typically occur on planar or curvilinear rupture surfaces or on relatively thin zones of intense
shear strain. When circular slope movements are involved, the potentially unstable mass can
be analysed by dividing it into slices with an assumed linear base. The forces and moments
acting on each slice can be calculated and summed to calculate the factor of safety for the
entire rupture surface.
The stability of a slope depends on the specific rupture surface that has the lowest FS,
known as the critical rupture surface. When the FS > 1, the slope is considered stable; when
FS is < 1, the slope is considered unstable; when the FS = 1 the slope is considered to be in
equilibrium. Although seeming relatively simple, limit equilibrium analyses require
considerable experience and judgment to select representative and appropriate sub-type of
movement (planar, circular rotational, and non-circular rotational), geological conditions
(stratigraphy, structural discontinuities and groundwater conditions) and geotechnical
properties (strength parameters of the rock or soil and their variability over the slope).
Discussions on the applications, limitations and assumptions of numerous limit equilibrium
methods are found in a number of publications such as Fredlund and Krahn (1977), Duncan
(1996), Hungr (1997), and Krahn (2003).
Although limit equilibrium methods are typically used for circular sliding types of
movements, in some instances, they can also be applied to the stability of soil and rock
masses controlled by one or several discontinuities, for example joints, faults, and clay seams.
In the simplest case, the limit equilibrium analysis of planar sliding consists of calculating the
shear strength along one sliding surface incorporating the influence of pore water pressure and
tension cracks, as appropriate (Wyllie and Mah, 2004). The general analytical solution for the
stability of a wedge formed by the intersection of two discontinuities, a tension crack and
sloping surfaces is summarised in Wyllie and Mah (2004). Similar methods have also been
proposed for toppling (cf. Goodman and Bray, 1976) and for more complex types of
movement such as bi-planar, ploughing, and buckling (Cavers, 1981; Hawley et al., 1986;
Alejano et al., 2011).
Some of the common, simpler limit equilibrium methods, that satisfy force or moment
equilibrium, include Bishop (1955) and Janbu (1973) whereas more rigorous methods, that
satisfy both force and moment equilibrium, include Morgenstern and Price (1965), Spencer
(1967) and Sarma (1973). Several methods extend the analysis from two-dimensional to
three-dimensional, for example Baligh and Azzouz (1975), Chen and Chameau (1982), Hungr
(1987, 1994), Hungr et al. (1989), and Hungr and Amann (2011). Recent efforts have been
made to integrate limit equilibrium methods into Geographic Information System (GIS)
environments, for example Reid and Brien (2006) and Grenon and Hadjigeorgiou (2010).

15

4.2.3. Numerical Modelling of Landslides


The application of numerical models to landslide investigations has increased considerably in
the last 10 years (Stead et al., 2006; Stead and Coggan, 2012). It is important that such
numerical models consider:
boundary conditions of the slope to be simulated (2-D or 3-D)
influence of structural geology on failure mechanism
material property and groundwater assumptions
in-situ stresses and landform evolution, and
parameter and model uncertainty.
4.2.3.1 Continuum Methods
The two principal methods of numerical modelling are the Finite Element Method (FEM) and
the Finite Difference Method (FDM). These methods are described by Jing and Hudson
(2002) and Itasca (2011). Both methods are well suited for landslide investigations in both
soil and rock. Numerous continuum codes can be used for slope stability analysis in soils with
a wide range of available constitutive models. The application of continuum codes to rock
slopes, however, requires the derivation of equivalent rock mass properties that reflect the
combined intact and discontinuity properties of the slope. Recent developments in continuum
numerical codes provide the engineer with the ability to:
determine the factor of safety using the shear strength reduction technique, SSR (Dawson
et al., 1999; Griffiths and Lane, 1999; Diederichs et al., 2007)
consider the influence of continuous, non-persistent and discrete fracture networks
(Hammah et al., 2009a; Sainsbury et al., 2008)
undertake probabilistic finite element analysis using the point-estimate method (Hammah
et al., 2009b)
investigate creep mechanisms in landslides (Grneng et al., 2010; Discenza et al., 2011)
undertake non-coupled or coupled groundwater-mechanical analysis of landslides
(Guglielmi et al., 2008; Cappa et al., 2004)
determine the importance of spatial property variation within a slope (Griffiths and
Marquez, 2007; Jefferies et al., 2008)
investigate the importance of landform evolution on landslides (Leith, 2012)
examine the influence of earthquake loading (described in Section 4.2.4)
Three-dimensional continuum codes are increasingly being applied to landslide studies.
Refer to Ambrosi and Crosta (2011), Chemenda et al. (2009) and Groneng et al. (2010) for
relevant applications and case studies. Sophisticated 3-D continuum codes exist with the
facility for coupled mechanical-groundwater and dynamic analyses. Three-dimensional
models require significantly more data and assumptions with important implications for
model and parameter uncertainty.
4.3.2.2 Discontinuum Methods
The two most common discontinuum numerical methods used in landslide simulations are the
Distinct Element Method (DEM) and the Discontinuous Deformation Analysis (DDA). These
methods are described by Jing and Stephansson (2007), Itasca (2011) and Khan (2010).
16

The DEM method differs from continuum methods in that it allows large scale
displacements on multiple discontinuity orientations, in addition to deformation of joint
bounded blocks. Input data required include the elastic properties (shear and bulk modulus)
and strength properties (cohesion, friction angle, tensile strength and dilation) for the rock
comprising the blocks. Elastic, elasto-plastic constitutive criteria are usually used for the
joint-bounded blocks, however, options exists to consider directional weakness through a
ubiquitous elasto-plastic model, strain softening and creep models. The constitutive criterion
for the discontinuity surfaces includes Coulomb slip, continuous yielding and an optional
Barton-Bandis criterion. Input properties controlling joint failure and joint displacement are
the joint cohesion, friction angle and tensile strength, and the joint normal and shear stiffness.
Further details on the available constitutive intact and joint constitutive criteria are presented
by Itasca (2011). DEM models of slopes include non-coupled or coupled mechanical
groundwater, mechanical thermal and dynamic models. Rock slope monitoring using
thermocouple and subsurface displacement instrumentation, in addition to mechanical-thermal
coupled DEM models, indicates that thermally induced displacements can play a role in
progressive rock slope failure (Watson et al., 2004; Gischig, 2011).
DEM models have been successfully used to simulate numerous major landslides including
Randa, Switzerland (Willenberg, 2004; Gischig, 2011), Aknes, Norway, (Kveldsvik, 2008),
Mt Eiger, Switzerland (Jaboyedoff et al., 2012) and Turtle Mountain, Canada (Froese et al.,
2012). Shear strength reduction, SSR, methods have also been integrated within DEM
methods (Diederichs et al., 2007).
Three dimensional DEM models are increasingly being used in landslide investigations, for
example the Downie Slide, Canada and the Beauregard instability, Italy (Kalenchuk, 2010).
This work is particularly noteworthy because the 3-D DEM models incorporate geostatistical
treatment of site investigation data and model constraint through long term slope monitoring.
Three-dimensional DEM models have also been used to investigate the influence of geometry
and geological structure on the kinematics of landslides, with reference to conceptual studies,
potential instability at Turtle Mountain, Alberta, and rock slides at Chehalis, and McCauley
Creek, BC (Brideau, 2010).
DDA methods although infrequently used in landslide analysis have been applied to
simulations of the 1963 Vaiont landslide, Italy (Sitar and MacLauglin, 1997) and the Aknes
rock slope, Norway (Kveldsvik, 2008).
4.2.3.3 Brittle Fracture Modelling and Landslides
Brittle fracture during rock slope failures has been recognized as an important factor in
landslides involving step-path failure surfaces and the failure of intact rock bridges. Methods
for modelling brittle fracture include the following.
Hybrid Finite-Discrete Element Method, FDEM: This method has found increasing
application in landslide analysis over the last decade (Eberhardt et al., 2004; Stead et al.,
2006). These codes allow the advantages of both continuum and discontinuum codes to be
maximized along with intact rock fracture. They incorporate fracture mechanics principles
and allow a transition from an intact rock/rock mass to landslide debris. A major advance in
FDEM models has been the incorporation of Discrete Fracture Networks using codes such as
Fracman (Golder Associates, 2012) within the FDEM models.

17

DEM Voronoi or damage models (Alzoubi, 2009): In this method the jointed slope rock
mass in a DEM model is discretized into polygons whose sides are given intact rock
properties. If the shear or tensile stresses in the rock slope exceed the strength properties of
the polygonal block boundaries then fractures form and propagate through the polygonal
mesh potentially resulting in slope failure.
RFPA method (Tang and Hudson, 2010): This is a modification of the FEM that has been
used to investigate fracturing during rock slope failure.
Particle Flow Codes, PFC2D/3D (Wang et al., 2003; Poisel and Preh, 2008a). In this
method the jointed rock slope mass is formed by particles (2-D) or spheres (3-D) that are
bonded together at their contacts. Joints can be included with the particle flow code model
allowing both sliding/separation on planes of weakness or intact rock fracture when induced
stresses exceed bond strength. PFC methods have been used to model both landslide
triggering and flow.
LatticeSpring Methods (Cundall and Damjanac, 2009): In this method, the slope rock
mass is represented by nodes and springs; the nodes replacing the particles in a PFC model
and the springs representing the bonds between particles. Intact rock facture in a jointed rock
slope is represented by breakage of the springs. With this method, the capability for both 2-D
and 3-D brittle fracture modelling exists.
4.2.4. Seismic Slope Analysis
Seismic slope stability analysis can be divided into pseudostatic limit equilibrium methods,
permanent-displacement methods and stress-deformation numerical modelling (Jibson, 2012).
The decision as to the appropriate method of analysis must consider whether the slope hazard
is due to liquefaction, strain softening or other failure mechanisms. A staged approach can be
used to determine an initial pseudostatic limit equilibrium factor of safety with a defined
seismic coefficient, k. Based on the results of this analysis, methods that consider slope
displacement along the slip surface, or pseudostatic limit equilibrium with slope displacement
based seismic coefficients, can be used. For critical high risk slopes or landslides where
considerable data have been collected, more complex analyses involving numerical stressstrain methods can be utilized. A useful flowchart for seismic slope analysis is presented in
APEGBC (2010, Figure 4.1).
4.2.4.1 Pseudostatic Limit Equilibrium
Pseudostatic limit equilibrium method of analysis is the simplest and most commonly
available. Earthquake loading is typically represented as a constant horizontal force, kW,
acting through the centre of gravity of the potential slope failure. The term W is the weight of
the soil/rock mass above the potential slip surface and k is the seismic coefficient, equal to a
fraction of the Peak Ground Acceleration (PGA).
Jibson (2012) discusses the limitations of this method including the assumption of a
constant one direction earthquake force, which can be both highly conservative and physically
incorrect because the peak acceleration acts for only short periods and acts both upslope and
downslope. The selection of the appropriate seismic coefficient, k, is also a perceived
limitation; many assume a value of k = 0.15, but it should be emphasised that most published

18

values of k are based on calibrations from very different slope environments to the landslide
being considered. Refer to Jibson (2012) for further details.
Jibson (2012) also notes that in some cases, pseudostatic analysis can be non-conservative,
for example where earthquakes induce high pore water pressures or materials experience a
significant reduction in peak shear strength during shaking. He states that although
pseudostatic limit equilibrium analysis is widely used it is being gradually replaced by
permanent displacement and numerical modelling methods.
Wyllie and Mah (2004) describe the applications of pseudostatic limit equilibrium methods
to seismic slope stability analysis of rock slope stability. Existing commercial codes for
analyzing planar, wedge, circular and non-circular failure in slopes routinely include
pseudostatic analysis (Rocscience, 2011). In addition to considering horizontal earthquake
loading, facility exists to include vertical and inclined seismic loads using appropriate seismic
coefficients. Jibson (2012) notes that it is usually acceptable to consider only a horizontal
seismic coefficient.
4.2.3.2 Permanent Displacement Analysis
The most common slope displacement method is the Newmark rigid block analysis proposed
by Newmark (1965). Recognizing that a yield acceleration exists, at which a rigid block
above a potential failure surface has a factor of safety less than one and moves, allows a more
realistic analysis than the pseudostatic limit equilibrium analysis. Those portions of the
earthquake strong motion record where the yield acceleration is exceeded are integrated twice
to determine the velocity of the rigid block and the cumulative Newmark displacement.
Although individual peaks in the acceleration result in a temporary factor of safety less than
one, it does not necessarily mean that global failure of the slope will occur.
Three principal types of Newmark analyses are used: the rigid block Newmark analysis;
decoupled analysis and coupled analysis.
Rigid block analysis: This is the most commonly used analysis, although it suffers from the
limitation that it assumes that the landslide does not deform internally.
Decoupled analysis: This analysis extends the rigid block analysis to consider internal
deformation. The decoupled analysis involves a two-stage process. The first stage involves a
dynamic response analysis of the slope with no assumed failure surface. An average
accelerationtime history for the slope mass above the potential failure surface is developed
through estimation of the accelerationtime histories at multiple points in the slope. The
average acceleration has been referred to variously as k or HEA, the horizontal equivalent
acceleration with peak values termed k max or MHEA, the maximum horizontal equivalent
acceleration, Jibson (2012). The second stage involves inputting the derived acceleration time
history into the rigid block analysis as before to find the permanent displacement. The site
response analysis may be undertaken using public domain codes such as SHAKE (Schnabel et
al., 1972).
A public domain Java code (Slope performance during an earthquake;
http://earthquake.usgs.gov/research/software/slope_perf.php) exists for undertaking both
Newmark rigid block and decoupled permanent displacement analysis (Jibson and Jibson,
2003).
Coupled analysis: In this analysis, the dynamic response and the permanent displacement
analyses are fully coupled so the influence of the sliding on the ground motions is allowed.
19

Both decoupled and coupled analyses require determination of the shear wave velocity of
the soil, the thickness of the potential landslide and the damping ratio. Jibson (2012) states
that the Newmark rigid block analysis is suitable for fairly shallow earthquake landslides
(which form 85% of earthquake-triggered landslides) but recommends coupled analysis for
larger, deeper landslides. Further details on Newmark analyses and its variants are provided
by Jibson (2012).
4.3.4.3 Stress-Strain Numerical Analysis
Continuum and discontinuum numerical stressstrain methods can be used to investigate the
influence of earthquakes on slope stability and landslide failure mechanisms. Such analyses
provide an improved understanding of the internal processes operative within a slope during
seismicity but demand increased data quantity and quality in addition to the development of
sophisticated constitutive models.
Soil Slopes: Seismic slope stability analyses of soil slopes are traditionally undertaken with
equivalent linear seismic methods which are used to model wave transmission in layered
media and dynamic soil-structure interaction. FDM models are increasingly used in the
dynamic analysis of soils and soil slopes and are fully nonlinear methods. A comparison of
the two methods is presented in Itasca (2011) and Naesgaard (2011). Important considerations
in dynamic modelling of soil slopes using FDM models are the dynamic input (either
acceleration, velocity, stress or force histories), boundary conditions (free field, viscous),
mechanical damping and choice of constitutive model. A detailed discussion of the theory and
procedure of dynamic FDM analysis is provided in Itasca (2011)
A considerable amount of experience has been gained in the use of the finite difference
codes in seismic slope stability analysis. Naesgaard (2011) provides a useful summary of
seismic slope stability analysis and available constitutive criteria. When considering seismic
analysis of soil slopes that may be subject to liquefaction or flow, the use of a specialized
constitutive criterion is required. The FDM code, FLAC (Itasca 2011) can use numerous
constitutive models to consider total stresses (UBCTOT), a hysteretic model for nonliquefiable clays/silt soils (UBCHYST), and coupled effective stress analysis in sands
(UBCSAND). Refer to Itasca (2011) and Naesgaard (2011) for further details of constitutive
models used in dynamic modelling of soil slopes.
Rock Slopes: Both continuum and discontinuum numerical methods can be used to
undertake seismic slope stability of rock slopes although it is not common due to the required
input data and complexity of the rock mass structure. Dynamic analysis can be undertaken
using both 2-D and 3-D continuum and discontinuum models, however 3-D dynamic analyses
are typically limited to critical rock slopes with significant data, such as dam abutment slopes.
Bourdeau et al (2004) provide a typical example of the use of 2-D FDM modelling in the
investigation of the factors influencing seismic amplification, an important factor in landslide
triggering during earthquakes. In dynamic analyses, an input source is applied at the base of
the FDM model which can be a synthetic acceleration waveform or an earthquake record. An
example of the application of two-dimensional continuum seismic slope stability analysis is
provided by Bozzano et al. (2011). Examples of the application of two-dimensional
discontinuum seismic slope stability analysis of landslides/rock slopes have been published by
Eberhardt and Stead (1998), Kveldsvik (2008), Bhasin and Kaynia (2004), Havenith et al.
(2002) and Moore et al. (2011; 2012).
20

Important considerations in simulating the effect of earthquakes include the choice of


boundary conditions (free field, viscous), damping parameters and the characteristics of the
source waveform relative to the mesh size. Guidelines are presented in Itasca (2011). Moore
et al. (2012) present a recent study using a DEM code to study the influence of compliant subvertical fractures (opened due to gravitational displacements) on spectral amplification. Their
analyses focus on the Randa rock slide and the Rawilhorn rock avalanche, both in
Switzerland, and show the effects of topographic amplification and a hitherto largely
neglected amplification due to pre-existing damage from previous instability and seismicallyinduced damage.
4.2.5. Landslide Runout Analysis
When a potential landslide source has been identified, a runout analysis may be required to
estimate hazard or risk downslope; to estimate input parameters, such as velocity and flow
depth, for treatment; or to estimate possible secondary effects, such as landslide-generated
waves or flooding caused by landslide dams.
The simplest form of runout analysis involves qualitative assessment of topography,
based on experience and judgment, to estimate the potential travel direction, distance and
velocity. Quantitative methods of analysis range from simple empirical-statistical
relationships to complex three-dimensional numerical modelling. The method used depends
largely on the type of landslide in question and the purpose and scope of the study. For
overview or preliminary assessments, empirical methods are often adequate; for detailed risk
assessment or treatment design, it is common to supplement empirical estimates with
numerical modelling. Some of the available methods, with references to more detailed sources
of information, are described below.
4.2.5.1. Empirical Methods
Topographical Analysis: The potential travel direction of a landslide can often be estimated
using topographic maps. If digital topographic data are available, GIS mapping software can
be useful, for example to identify topographic fall lines. Potential velocities and runout
distances are more difficult to estimate qualitatively, but can be constrained by records of
similar landslides in the area using professional experience and judgment.
i) Volume-fahrbschung Relationships: There is typically an inverse relationship between
landslide volume and fahrbschung, the angle of the line connecting the crest of the
landslide with the toe of the deposit, measured along the approximate centre-line of motion.
Fahrbschung is also known as the angle of reach. Regression equations have been proposed
by several workers (Scheidegger, 1973; Li, 1983; Nicoletti and Sorriso-Valvo, 1991;
Corominas, 1996; Legros, 2002; Hunter and Fell, 2003) and typically take the following form:
H
log10
L

a log10 V + b
=

Equation 4-5

where H is the elevation difference between the crest of the source and the toe of the
deposit, L is the horizontal distance between the crest and toe along the centre-line of
movement, V is the landslide volume, and a and b are coefficients. If such relationships exist,
21

or can be established for a specific region or landslide type, they can be used to estimate the
travel distance.
ii) Energy Grade Lines Relationships: These relationships are similar to volumefahrbschung relationships, but are based on a line drawn between the centres of mass of the
source and the deposit. The inclination of this line, also known as the travel angle, represents
the average bulk basal friction angle of the landslide (Hungr, 1981). Using energy grade line
principles established in hydraulic engineering (for example Roberson and Crowe, 1993),
flow velocities at any point along the path can be estimated based on the elevation difference
between the energy grade line and the flow path. The challenge with this approach is that it is
difficult to accurately estimate the position of the centre of mass in the source and deposit
zones. As a result, unlike volume-fahrbschung relationships, statistical correlations between
landslide volume and travel angle are uncommon. The assumption of a constant bulk basal
friction angle also tends to overestimate flow velocities.
iii) Volume-area Relationships: These relationships involve statistical correlations
between landslide volume and depositional area. Such relationships have been proposed by Li
(1983), Hungr (1990), Iverson et al. (1998), and Legros (2002). The regression equations
typically take the following form:
A = cV

2
3

Equation 4-6

where A is the planimetric depositional area, and V is the landslide volume. Values of the
coefficient c can range from approximately 10 for rock avalanches to 200 for lahars (Hungr,
1990; Griswold, 2004). Volume-area relationships are the basis for the US Geological
Surveys GIS-based mapping program LAHARZ (Iverson et al., 1998; Griswold, 2004). Note
that volume-area relationships only provide estimates of the depositional area and must be
used in combination with other methods to estimate where the deposit may be located along
the path. LAHARZ requires the starting point of deposition to be specified by the user.
iv) Volume Balance: Models based on the volume balance of material entrained and
deposited along the debris flow travel path can be used to estimate runout distances,
distribution of erosion, and deposit depths (Cannon, 1993; Fannin and Wise, 2001). Such
models require the input of detailed path morphology and must be locally calibrated. Volume
balance is the basis for the web-based model UBCDFLOW (Fannin and Wise, 2001).
v) Rock Fall Shadow Angle: For rock falls, Evans and Hungr (1993) proposed a runout
mapping method based on the concept of shadow angle, defined as the angle between the
apex of the talus slope and the distal margin of rock fall beyond the toe of the talus slope.
They examined 16 talus slopes in southwestern British Columbia, which yielded a minimum
rock fall shadow angle of 27.5. Local calibration of the method is required.
vi) Other Empirical Methods: Other methods have been proposed to estimate landslide
runout distances and velocities. Hsu (1975) proposed a correlation between landslide volume
and excess travel distance, which is the distance beyond what would be predicted using a
normal average bulk basal friction angle of 32. Rickenmann (1999) presented statistical
correlations to estimate debris flow peak velocity and discharge.
4.2.5.2. Analytical Methods
i) Forced Vortex Equation: This equation for superelevation was originally formulated for
22

open channel water flow and was adopted for snow avalanches (Mears, 1981) and debris
flows (Hungr et al., 1984). It is most commonly used to back-calculate debris flow velocities
based on post-event observations of trimlines in channel bends, but it can also be applied to
deflection berm design (Hungr et al., 1984). The following equation is based on a lateral
momentum balance at the point of maximum superelevation:
v2
H =
B
Equation 4-7
Rg
where H is the elevation difference between trimlines on the inside and outside of the
channel or gully bend, B is the surface width of the flow, v is the mean flow velocity, R is the
mean radius of curvature in the bend and g is the acceleration due to gravity.
Leading Edge Model: This model was originally developed to estimate debris flow
runout (Takahashi and Yoshida, 1979). It was modified by Hungr et al. (1984) for runup
against adverse slopes, such as terminal berms, and was subsequently applied to snow
avalanches (Hungr and McClung, 1987; McClung and Mears, 1995). The following equations
are based on a longitudinal momentum balance at the point of maximum runup:
V2
H=
sin
G
Equation 4-8
gh cos
V v0 cos ( 0 ) 1 + 0 2 0
=
2v0

G g ( cos sin )
=

Equation 4-9
Equation 4-10

where H is the runup height, is the adverse slope angle (negative), 0 is the approach
slope angle (positive), v0 is the approach velocity, h0 is the approach flow depth and is the
basal friction coefficient.
4.2.5.3. Physical Models
The first large debris flow deflection works in British Columbia were designed with the help
of 1:240 scale model tests with a bentonite slurry (Nasmith and Mercer, 1979). The
requirements for dynamic similitude in physical modelling of debris flows, however, are very
complex (Iverson, 1997), making physical model testing relatively costly and therefore
infrequently used.
4.2.5.4. Numerical Methods
Numerical models simulate landslide movement by time-step solutions of a governing set of
equations of motion. Such models allow unique geometry and local material characteristics to
be accounted for explicitly, and provide estimates of velocities and flow depths at different
points along the path. Both discontinuum and continuum numerical models have been
developed to simulate landslides in 2-D and 3-D. For both, calibration of model input
parameters is required.
i) Discontinuum Models: These models treat landslides as solid bodies, or assemblages of
solid bodies, with up to four different movement types: falling, bouncing, rolling and sliding.
23

The simplest models treat the landslide as a single body or lumped mass. Lumped mass
sliding block models, which are related to the energy grade line methods mentioned above,
have been applied to landslides and are routinely applied to snow avalanches (cf. Perla et al.,
1980). Lumped mass bouncing ball (fall-collision-rebound) models are routinely used for
rockfall simulation. Available rock fall models include RocFall (Stevens, 1998), CRSP
(Pfeiffer and Higgins, 1990) and STONE (Guzzetti et al., 2002). Calibration is required to
constrain the restitution and friction coefficients that are used in these types of models.
ii) Discrete/distinct element models: These models are a multi-particle extension of the
lumped-mass approach described above. In such models, landslides are modelled as a
collection of individual particles that interact with each other and with the ground surface by
falling, bouncing, rolling, and sliding to simulate large-scale deformations. The particles can
be of a variety of shapes and sizes, and a variety of inter-particle and particle-bed contact
relationships can be modelled. The Particle Flow Code (PFC) model, developed by Cundall
and Strack (1979), uses circular (in 2-D) or spherical (in 3-D) particles that can be bonded
together or broken apart under certain conditions (Poisel and Roth, 2004). PFC has been used
by several landslide professionals to simulate landslides (Calvetti et al., 2000; Gonzlez et al.,
2003; Pirulli et al., 2003; Poisel and Preh, 2008b).
iii) Continuum Models: These models are based on established depth-averaged (shallow
flow) hydrodynamic models of landslides. In contrast to the solid body approach of
discontinuum models, continuum models treat the landslide as a fluid-like material. Flow
spreading in depth-averaged continuum models is governed by internal stress gradients, and
basal shear stresses provide resistance to forward motion. Different rheological models (for
example, frictional, Voellmy and Bingham) can be used to estimate the basal shear stress at a
given time and location within the flow. A number of different continuum models are
currently in development around the world (Hungr et al., 2007). Available models include
DAN (Hungr, 1995), TITAN2D (Pitman et al., 2003) and FLO-2D (FLO-2D Software, 2007).
Extensive calibration work has been carried out using DAN (Hungr and Evans, 1996; Ayotte
and Hungr, 2000; Revellino et al., 2004; Pirulli, 2005; McKinnon, 2010) and a 3D version of
the model, DAN3D, has been developed (McDougall et al., 2006).
4.2.5.4.1 Associated Techniques
i) Model Calibration: Calibration is required to constrain numerical model input
parameter values. Typically, rheological parameters are constrained by systematic adjustment
through trial-and-error or statistical back-analysis of full-scale prototype events. On a caseby-case basis, calibration is achieved by matching the simulated travel distance, deposit
distribution, flow velocities, and travel times to those of the prototype landslide, usually
through subjective visual assessment. More advanced calibration techniques have recently
been proposed by Galas et al. (2007), Cepeda et al. (2010) and McKinnon (2010).
ii) Probabilistic Mapping: Runout predictions based only on a single set of input
parameter values can be misleading unless they are placed in the proper context. Depending
on the level of required conservatism, the results may represent a workers best guess, with a
subjective probability of exceedance of 50%, or they may represent a more conservative
estimate, with an exceedance probability of 10% or less. Many of the methods outlined above
can be applied in a probabilistic framework that addresses this issue and provides results that
are suited to quantitative landslide risk assessment (McDougall, 2012).
24

One type of probabilistic approach is to determine the best-fit results for each individual
case, as described above, and plot the results as a histogram, which can be fitted to a
probability density function. The probability density function can then be used to assign
exceedance probabilities to parametric model runs, or to define input value probabilities for
use in Monte Carlo-style predictive runs (Dalbey et al., 2008). This approach is commonly
used in rockfall mapping.
Another approach is to determine the best-fit results for each group of similar cases as a
whole and then attempt to quantify the resulting variance. The results can be used to place
confidence bounds on predictions (Schilling et al., 2008), which can be translated into explicit
exceedance probabilities.

5.0 MONITORING
In general, known landslides that are a risk to public infrastructure, private property, and
personal safety should be monitored in order to minimize the potential risk. The monitoring
program, whether it is simple or complex, must be designed based on the geological, physical,
and environmental conditions of both the landslide hazard and the elements at risk. As the
definition of landslides includes a very broad range of natural phenomena, the range of
techniques commonly used to monitor their stability is also quite broad.
There are two paramount reasons for monitoring landslides:
To understand the dynamics of the mass movement to facilitate the assessment of
the hazard;
To assess the degree of instability through detection of changes which enable the
development of accurate early warning systems and precautionary measures.
5.1.

Slope Engineering

Landslides occur naturally, as a result of human construction, and in engineered slopes. The
nature of the problem, however, remains consistent: unstable slopes that can negatively affect
personal safety or property must be monitored to ensure minimal risk. Known landslides can
be separated into two categories, those in which support methods can be implemented to
reduce or eliminate the hazards, and those in which monitoring must be performed to reduce
the effect of the hazard through implementation of early warning systems. In both instances
monitoring is used to delineate the hazard; however, in the first case, intensive monitoring is
generally reduced once the hazard has been reduced or eliminated through support/remedial
measures.
Landslides occurring in natural terrain include (but are not limited to) debris flows, rock
slides, rockfalls, rock avalanches, quick clay/sensitive clay failures, and traditional landslides.
The occurrence of natural slope landslide can occur in any geological, geomorphological, or
physical setting. This category also includes natural landslides which develop due to
engineered constructions. Landslides occurring in engineered slopes can include structures
such as dams, embankments, and retaining walls.
Each of the monitoring methods discussed in the following section can be successfully
applied to both natural and engineered slopes.
25

5.2.

Monitoring Methods

Monitoring methods for slope instability include a wide variety of instruments and techniques.
Commonly slopes are monitored for displacement, differential movement, hydrological
changes, strain, and ground temperature, whereas the physical environment is monitored for
temperature and precipitation. Table 5 contains essential remote sensing based techniques for
landslide monitoring; they are arranged according to the method used to collect or generate
the data. Table 6 presents in situ monitoring techniques. A broader discussion of each major
category is contained in Sections 5.2.1-5.2.4.
Table 5. Fundamental remote sensing based landslide monitoring methods.
Method
Terrestrial
Lidar
Photogrammetry
Robotic Total Station
Targets
GB-InSAR
Digital Image Correlation

Airborne
Lidar
Multispectral Imaging

Spaceborne
Very High Resolution
Imaging
InSAR

Application

Accuracy

Full field displacement/discrete


failure
Full field displacement/discrete
failure
Discrete location displacement

mm dm (~with distance)

Full field displacement


Full field displacement

mm
Mm

Full field displacement


Environmental changes

cm m (~with distance)
cm m

Physical and environmental changes

dm m

Full field displacement

mm cm

mm dm (~with distance)
mm cm (~ with distance)

Table 6. Fundamental in situ based landslide monitoring methods.


Instrument Type

Deformations/Displacements
Inclinometers (Vertical &
Horizontal)
ShapeAccelArray (SAA)
Extensometers
Settlement Gauges
Tilt Sensors & Tilt Beams
Pendulums & Plumblines
Fibre optics BOTDR

Application

Accuracy/
Comments

Monitor horizontal or vertical offsets


within ground mass
Monitor settlement and deflection
Monitor dilation or contraction across
geologic features
Monitor vertical settlement or uplift
Monitor tilt changes at discrete locations
Monitor horizontal offsets from vertical
Monitor 3D shear strain

1-2 mm/30 m

26

Sub 1 mm
Sub 1 mm
1-2 mm
0.1 mm/m
0.5 mm
0.5 mm

Differential GPS

Monitor horizontal and vertical changes to


surface positions
Monitor the progression shear strain
perpendicular to drillhole axis

5-10 cm vertical and


horizontal
1mm/m

Monitor water levels and/or water pressures


Monitor seepage rates and changes
Monitor groundwater temperature at discreet
locations
Monitor groundwater temperature at multiple
locations along an axis
Monitor various chemical and physical
properties of groundwater to identify changes.
pH, salinity, etc.

0.5-1 mm
0.1-1 litre/min
0.1 degree C

Monitor temperature
Monitor amount and rate that site receives
water
Monitor barometric pressures at the site
Monitor surface humidity at the site
Monitor wind
Monitor surface evaporation conditions
Monitor hours of sunlight
Monitor snow accumulation and rate at which
spring melt occurs

0.1 degree C
0.1 mm absolute

Triggered device that will record major seismic


events that occur at a remote site
Major seismic recording device. Would only
be required under special circumstances

will only record


events > M 6.0
can capture very
small seismic events

Acoustic Emission

Micro-particle ground motion

unknown

GNNS GeoCube

Monitor horizontal or vertical movement

sub 1 mm

Time Domain Reflectometry

Groundwater/Seepage
Piezometers
Weirs & Fumes
Temperature Probes
Thermister Strings
Water Chemistry

Environmental/Climatic
Air Temperature
Precipitation
Barometric Pressure
Humidity
Wind; Direction & Speed
Evaporation
Sun Light Hours
Snow Pillows

Seismic Activity
Strong Motion Accelerometer
(SMA)
Seismograph

0.1 degree C
various

0.01 KPa
various
various
various
0.1 hrs/day
cm/day

Upcoming Methods

5.2.1. Remote Terrestrial Techniques


High resolution terrestrial remote sensing techniques enable engineers to identify instabilities,
detect movement, and characterize slope deformation. The most commonly employed and
published technique is LiDAR Light Detection And Ranging (colloquially referred to as
laser scanning). Other well established techniques include: terrestrial digital photogrammetry,
robotic total stations, and ground-based interferometric synthetic aperture radar (GB-InSAR).
These techniques present interesting advantages in comparison with existing monitoring
methods; their application frequently leads to a better understanding of landslide failure
mechanisms, rock fall rate and spatial distribution, as well as pre-failure deformation.
27

The principle of each of these instruments is the ability to generate data depicting the
physical location of the landslide area at a given point in time. The ability to use one of these
methods for monitoring displacement within a landslide requires a temporal repetition of data
collection. Different data analysis methods can then be used to quantify slope movements,
based on the comparison between sequential high resolution 3D datasets. If there are
significant morphological variations between successive acquisitions (e.g., rockfall), simple
change detection procedures can be applied. In the case of relatively slow and homogeneous
displacements with no significant morphologic changes, it is possible to characterize more
completely the displacement and deformation of a moving mass. Figure 2 illustrates the
scanning on an active landslide zone in British Columbia above a railway track. LiDARderived information is being used to assess the kinematic stability and displacements of the
slope face.

Figure 2. LiDAR scanning in western British Colombia to assess kinematic stability and monitor
displacements.

The use of terrestrial digital remote sensing techniques for the characterization and
monitoring of landslides may lead to new perspectives in their understanding, stability
assessment and mitigation. The literature in this field is significant, Table 8 includes a few
key papers on the various methods.
5.2.2. Airborne Techniques
Airborne monitoring of landslides is typically conducted using LiDAR, photogrammetry, or
28

multispectral imaging techniques. The techniques are very different in terms of capability.
Airborne LiDAR Scanning (ALS) data, collected from an airplane or helicopter, is most
commonly used to delineate terrain surface models and identify large unstable masses.
Sequential ALS data can be used to map differential change on a slope over time which can
aid in the understanding of active slope processed (Lato at el., 2014). Oblique Helicopter
Photogrammetry (OHP) is a developing method of collecting photos for 3D surface data
generation. Optimally, OHP data is collected at a range of 100-500 m from the rock face
using high quality handheld camera equipment. The view angle of the camera with respect to
the slope of interest can be adjusted in real-time by the photographer. Data collection is
therefore only restricted to the flight path of the helicopter operator; as a result occluded
regions can be minimized through flight planning (Lato et al., 2015).
5.2.3. Spaceborne Techniques
Current state-of-the-art in real-time monitoring of active slopes developed for early warning
of landslides is very expensive. Satellite radar interferometry is used to complement real-time
monitoring such as GPS, TLS, in-situ field measurements, and others. Interferometric
Synthetic Aperture Radar (InSAR) techniques are being used to measure millimetre
displacement on slow-moving landslides. Key references are included in Table 8.
An interferometric phase image (interferogram) represents the phase differences between
the backscatter signals in two or more SAR images obtained from similar positions in space.
In case of spaceborne SAR, the images are acquired from repeat pass orbits. Once the
topographic phase is removed, the phase differences between two repeat-pass images are the
result of changes in topography, changes in the line-of-sight distance (range) to the radar due
to displacement of the surface and change in the atmospheric conditions between scenes. For
non-moving targets, the phase differences can be converted into a digital elevation model.
InSAR techniques are being used to monitor gradual landslide motion under specific
conditions, provided coherence is maintained over the respective orbit cycle. Landslide
movements are measured in millimetres to centimetres per orbit cycle of the radar satellite.
This orbit cycle can range from 44 days for ALOS, 24 days for RADARSAT 2, 11 days for
TerraSAR X, and 2.5 days for Cosmo-SkyMed.
Reliable measurements of surface displacement can be achieved under specific
conditions. These include using radar image pairs or numerous scenes (more than 25), with
similar viewing geometries, short perpendicular baselines (less than 100 m), short time
intervals between acquisitions, and correcting for the effect of topography and atmospheric
conditions. The InSAR deformation maps or profiles provide linear motion along the line-ofsight. However, under some geotechnical assumptions assessments can be refined to vertical
and horizontal components. In the early stages of the research, 3D-motion using ascending
and descending orbits and additional viewing geometries is produced.
InSAR processing techniques include differential InSAR, PSInSARTM /CTM TM, and
more recently Squeeze SAR TM. The differential InSAR uses only two or three scenes or
orbit cycles and provides a more general and simplified snapshot of the deformation activity
of the landslide. It is a powerful tool to measure displacements because it offers a synoptic
view of the landslide. The more detailed point targets or corner reflector data using
PSInSARTM/ CTM TM or Squeeze SAR TM processing techniques are more complex. The
InSAR deformation profiles for both point and distributed targets show the spatial and
29

temporal heterogeneities of terrain movements and assist in defining the parameters


controlling the dynamics of low-velocity landslides. This simple guideline on InSAR
monitoring of Canadian landslides should be revised as more case studies are documented.
The uses of InSAR monitoring are rapidly increasing with the availability of more frequent
satellite revisits from our future RADARSAT Constellation Mission (RCM), combined with
improved InSAR processing techniques taking advantage of several viewing geometries,
distributed targets and smaller corner reflectors.
5.2.4. In Situ Techniques
In situ monitoring techniques are generally employed as part of the surface and sub-surface
investigations that take place on a landside site. Despite the quality and quantity of the
detailed information that can be collected and analysed by remote sensing methods, at some
point, decisions have to be made regarding whether the investigations need to shift toward
intrusive in situ monitoring techniques. Remote sensing information, coupled with surface
investigations and geological mapping, will enhance the understanding of the physical
characteristics and the genesis of the landslide mass to a certain point. This work will always
result in many questions regarding the geometry and mechanism of the slope failure and the
rate of the actual ongoing ground movements.
Drilling investigations and excavations are the usual intrusive methods that are employed
to recover detailed information regarding the slope geometry, the failure mechanism of the
slope and the current rate of movement. It is noted that these intrusive methods can be very
costly and therefore require compelling justification for the expenditures. Usual justifications
are for either commercial reasons or in the interest of public safety. Around the world today,
most major intrusive investigations and in situ monitoring programs that are carried out on
landslides are for the protection of commercial interests and to mitigate the related financial
liabilities.
Drilling and excavations (either surface or underground) provide many opportunities for
the inclusion of in situ monitoring methods, which will provide key information on the
behaviour and physical nature of the landslide mass. These in situ monitoring methods are
summarized in Table 6.
The in situ monitoring instrument types which are most commonly used and provide the
key information for landslide monitoring are Inclinometers and Extensometers (under
Deformation / Displacement Instruments) and Piezometers and Weirs and Flumes (under
Groundwater / Seepage Instruments). The roles of these key in situ monitoring instruments
are summarized below.
Deformation/Displacement:
Inclinometers: The installation of inclinometer casings in near vertical drillholes
are used to accurately measure the magnitude and time progression of deformation
zones that are occurring near perpendicular to the drillhole alignment. This
monitoring technique can also be carried out in near horizontal drillholes.
Inclinometers represent a very precise monitoring technique which is used
extensively in landslide masses to determine movement directions, magnitude of
movements and movement rates.
Extensometers: There are many variations of extensometer type instruments,
which are intended to measure convergence or divergence on an approximate
30

straight line between two end points. Extensometers are designed for many
applications on surface or within drillholes. Readings can usually be done by a
manual method, but are easily adapted for automated electronic readings.
Extensometer instruments can be designed to provide extremely precise data. They
are intended to measure extension or contraction movements, which are occurring
between the end points either within or on a landslide mass.
Groundwater/Seepage:
Piezometers: There are many variations of piezometer type instruments. These
instruments are intended to measure water levels within a slope or a landslide
mass, which is equated to the water pressure acting on the slide mass slip plane.
This piezometric information is critical to understanding the mechanism for a
slope failure and for determination of the current stability of a slide mass.
Standpipe piezometers are the simplest and most common water level measuring
instrument, but there are many different variations which provide extremely
accurate water pressure data.
Weirs and Flumes: There are many different variations on weirs and flumes which
are used to monitor seepage flow rates either entering or exiting a slide mass.
Naturally occurring seepage or seepage from excavations or drillholes are usually
routed to a collection point where measurement can be carried out accurately in a
weir or flume. Data are also required on the water input into the slope due to
precipitation (Environmental/Climatic). The accurate measurement of seepage
input and output flows is commonly used as input data for flow modelling of a
landslide slope mass. This modelling may indicate critical periods within the
annual cycle when slope stability is reduced and movement may be renewed.
5.3.

Landslide Warning Systems

Landslide warning systems often serve as an interim risk management measure in absence of,
or to complement, engineered structures. Warning systems are particularly helpful in
situations where significant landslide risks have been identified either regionally or sitespecifically, and funds are currently not available to address all sites that exceed a limit of
tolerable risk. In those cases, landslide warning may allow residents or workers to evacuate at
a critical time when a landslide is likely to occur.
Landslide warning can be based on deformation measurements and warning levels may
be defined when such deformations reach specified thresholds. Alternatively, because most
landslides are associated with some hydroclimatic factors, a combination of such factors could
be linked to specified warning levels. Figure 3 displays a landslide monitoring station
equipped with an extensometer, gravity-referenced inclinometer, and temperature sensor. The
system is powered by a rechargeable battery via a solar panel, and data recorded by the
system as well as the health of the system are checked via cellular connection.

31

Figure 3. In situ measurement system located in the tension crack of an unstable rockmass
800 meters above a small community. As shown in inset figures (b) and (c) the monitoring
unit is connected to a solar panel and control system to report measurements via cellular
uplink. The measurements taken (c) are extension and inclination across the crack, as well as
air temperature.
5.3.1. Early warning systems
Reliable prediction of landslide occurrence based on climatic thresholds has been
accomplished in numerous countries and has been described in many publications, some of
which are included in Table 7. A large number of researchers have used rainfall intensityduration curves and plotted rainfall data from landslide-producing and non-producing storms.
Envelopes or separating lines were drawn to extract the intensity-duration data that allow
classification of a storm to be either debris-flow producing or not.
An Example: The District of North Vancouver in collaboration with BGC Engineering
developed Canadas first operating real-time debris flow warning system, which operated
during the rainy seasons of 2009/2010 and 2010/2011. The system is based on discriminant
function analyses of 20 hydrometric input variables consisting of prior rainfall and storm
rainfall intensities for a total of 63 storms. Of these, 27 resulted in shallow landslides and
subsequent debris flows, while 36 storms were sampled that did not result in debris flows. The
discriminant function analysis identified as the three most significant variables the 4-week
antecedent rainfall, the 2-day antecedent rainfall, and the 48-hr rainfall intensity during the
storm. The resulting classification functions provide a scale representing the likelihood of
32

debris flow initiation. Several complexities were added to render the classification functions
into a usable and defensible warning system. Back-calculation of the models 21-year record
confirmed that 75% of all debris flows in the past 20 years would have occurred during
warning- or severe warning levels. Antecedent rainfall is automatically calculated as a sliding
time window for the 4-week and 2-day periods every hour. The predicted 48-hr storm rainfall
data are provided by the Fluid Dynamics Centre at the University of British Columbia and are
updated every hour as rainfall is recorded during a given storm. The warning system
differentiates five different stages: No Watch, Watch Level 1 (the warning level is unlikely to
be reached), Watch Level 2 (the warning level is likely to be reached), Warning and Severe
Warning.
5.3.2. Technical considerations and data connectivity
In general, a typical warning system will consist of four major elements that collectively will
provide a total system approach to risk management. These elements are: data recording, data
transfer, data review, and warning communication. Key references to these topics are included
in Table 8.
The ability to notify any change in alarm status on a warning system will be dependent on
the successful implementation of these elements, all of which are susceptible to failure for
various reasons. For instance, equipment failure might compromise the data collection and
data transfer elements; and the lack of qualified people generally constitutes a typical source
of trouble for the data review element. On the other hand, the warning communication
element can be severely affected if an emergency response protocol that clearly defines roles
and responsibilities in the event of an incident is not available.
Therefore, for any warning system, it is important to take the necessary steps to guarantee
that the system will function as designed on a near-continuous basis. Thus, the overall
approach that must be taken when designing and operating an early warning system is to
create a series of tasks that will ensure adequate functioning of each of the warning system
elements (Table 7).
5.4.

Summary and Limitations

The science of landslide monitoring is diverse and complicated. The numbers of options are
immense. Each technique is accompanied by a long list of limitations, sources of errors, and a
subject matter expert who can demonstrate its successful implementation on various projects.
When designing a monitoring program and deciding which technique(s) to use, it is critical to
first establish what parameters are essential to monitor and what resolution and accuracy are
required. These decisions are generally determined through analysis of existing knowledge of
the landslide with respect to its size, failure mode, activity level, and elements at risk.
The primary internal limitation of any monitoring systems is that it can only produce data
in the manner and/or spatial region which it is programmed to collect. Because the
determination of stability (or hazard level) can only be accessed through analysis of these
data, it is critical that the monitoring system be specifically designed for each potential mode
of instability. As well, each instrument or technique will have associated sources of error and
inaccuracies; this is the responsibility of the engineer/geoscientist to understand when
designing a monitoring system.
33

The cost of equipment, installation, and data analysis is often the largest external
limitation of any monitoring program. This, however, must be dealt with on a case by case
basis, and engineering judgement and expert opinion can help in determining the best
approach for the most successful results.
Table 7. Suggested tasks required to provide an effective warning system.
Element

Task

Data recording

Design a sensor network that provides complementary types of instruments with varying
sensitivities to movement and climatic influences
Perform regular system health checks
Develop well-defined plan for maintenance and repair to reduce potential downtime for the
system

Data transfer

Provide a main communication link that gives access to data and at least one secondary
(back up) link

Data review

Develop a protocol to ensure data are reviewed on a specified frequency and in a repeatable
manner (weekly review for early detection of increasing acceleration; yearly review for
long-term trend analysis)
Develop kinematic model for unstable mass and update as monitoring data become
available
Establish thresholds for alarm triggers (absolute and velocity-based)

Warning
communication

Develop set of rules for notification of increased alert level (Emergency Response
Protocol-ERP)
Run orientation sessions (training) and drills (testing) for the ERP
Have debriefing session after any alert incident

In addition, it should be noted that by initiating a monitoring program, there may be an


implied responsibility for the long-term continuation of the monitoring program in the eyes of
the public or involved government agencies. Therefore, caution is recommended to any party
intent on starting an instrumentation and monitoring program at a landslide site. When
establishing a monitoring program, the owner should consider two things. Firstly, determine
and establish up front, with all external parties, the circumstances and conditions by which the
monitoring program will be shut down or discontinued. And secondly, the party must be
careful to never commit to a monitoring program which cannot be sustained for the long term.
Complex monitoring programs carry major responsibilities, which are both expensive and
require a highly trained and dedicated professional workforce. Table 8 identifies important
publications that explain various remote sensing techniques and their application for mapping
terrain and monitoring slope activity.

34

Table 8. Key references and case studies for landslide monitoring.


Key reference
Remote Sensing
Pedrazzini et al., 2011

Abellan et al., 2014


Eberhardt et al., 2008
Lato et al., 2011
Lato et al., 2014
Bateson et al., 2015
Singhroy, 2008
Kromer et al. 2015

In Situ Monitoring
ICOLD, 1989

Baker, 1991
Moore et al., 1991

Early warning Systems


and Data Connectivity
Wieczorek and Glade, 2005
Guzzetti et al., 2005
Baum and Godt, 2010
Jakob et al., 2011
Froese and Moreno, 2014

Topic

Integration of terrestrial remote sensing-based discontinuity characterization and


slope monitoring into the hazard assessment analysis and warning system
implemented within the risk management procedure for a quarry
Review of the applications of terrestrial LiDAR for slope stability assessment and
monitoring
Demonstration of the capability of monitoring active landslides with robotic total
stations
Review and recommendations concerning the integration of LiDAR-based
measurements into rockfall hazard management systems
Review of airborne LiDAR, terrestrial LiDAR, and terrestrial Photogrammetry for
mapping change in active environments.
Application of the ISBAS algorithm for monitoring slope deformation
Satellite remote sensing applications for landslide detection and monitoring
Mapping millimetre-scale ground deformation along transportation corridors using
terrestrial LiDAR for precursor failure movements

International perspective on the state of the art regarding the in situ monitoring and
data acquisition for major structures. Provided information is directly applicable to
landslide and slope stability problems.
Details of the investigations and in situ instrumentation installations carried out for
the geotechnical monitoring of a major slope instability.
Details of how in situ instrumentation and near real-time monitoring can be used to
reduce risks due to potential slope failures.

Climatic factors influencing triggering of debris flows

Evaluation of Flood and Landslide Risk to the Population of Italy


Early warning of rainfall-induced shallow landslides and debris flows in the USA
Regional real-time debris-flow warning system for the District of North
Vancouver, Canada
Structure and components for the emergency response and warning system on
Turtle Mountain

6.0 SLOPE TREATMENT


6.1.

Introduction

Slope treatment is required in areas where a risk evaluation has determined that it is
unacceptable to leave a slope in its current condition. Treatments can be applied to both preexisting landslides and to the prevention of future landslides. Mitigation involves treatment to
reduce the seriousness or impact of the landslide, while remediation involves treatment to
35

reverse the damage caused by the landslide. This section will provide various options
available for slope treatment under different geological and environmental conditions.
6.2 Treatment Considerations
Any treatment program will have to be consistent with the objectives of the study. The study
area will have to be defined and the appropriate level of effort applied to the treatment
program will depend on the consequences of failure. Linear projects such as transportation
routes will typically involve a larger study area and will have different treatment options than
a localized structure.
Many factors must be considered in determining appropriate treatment options for
unstable slopes. The variability of soil conditions, environmental factors and economic
constraints results in a unique approach for almost every landslide. The best possible data
gathering and sound engineering judgement must be used to arrive at a safe and economical
solution to the problem.
Table 9 provides a summary of the more common treatment methods for unstable slopes.
The methods are divided into active and passive classifications. As shown in the table, there
are numerous options available and often a combination of options is selected for specific site
conditions. Treatment options will be affected by the local geology, the scale of the problem,
the consequences of failure, the economic situation and the resources available.
6.2.

Passive Methods

Passive methods are methods that do not involve physical stabilization of the slide mass but
rather attempt to avoid the problem. This is often a good approach during planning stages of a
project. Once design and development has begun, relocation or alteration of the design can
become costly and less desirable for the project stakeholders. The unstable areas must be
identified during the investigation to allow avoidance options to be determined.
Some regulatory agencies have developed guidelines or standards for dealing with
landslide prone areas. These documents typically prescribe methods of analysis and the levels
of deformation that are deemed acceptable for certain types of developments (typically
residential areas). For example, in British Columbia, the Guidelines for Legislated Landslide
Assessments for Proposed Residential Developments produced by the Association of
Professional Engineers and Geoscientists of B.C (Gerath et al., 2010)
Passive methods are also used where ongoing movements are small and the trigger
mechanism is understood. In this case the acceptance of the risk and the implementation of
monitoring can be used. For example, at the Gardiner Dam in Saskatchewan filling of the
reservoirs activates small movements annually (Rahman and Kilgour, 2000). These
movements are small, understood and monitored. The following sections summarize various
alternative passive methods to slope treatment.
6.2.1. Risk Assessment
6.2.1.1. Accepting the Risk
Risk based principles can be used to conduct probabilistic assessments of slope reliability.
36

Based on the results of such studies, it may be possible to accept the risk of landslide
occurrence and therefore leave things as they are. In some cases, small movements may be
ongoing, however, if they are understood, it may be possible to accept the movement. In some
locations, risk tolerance may be defined by local regulators. However, in other areas risk
acceptance levels may not be possible to define.
Table 9. Summary of treatment methods for unstable slopes.
CLASS

PROCEDURE
Risk assessment

Passive methods

Avoidance
Monitoring and
warning

METHOD

DETAILS

nothing required
regulatory methods
non-regulatory methods
avoid unstable ground

risk is acceptable
covenants, restrictions
land swaps, education
relocate structure, bridge over

survey hubs, GPS, LiDAR inclinometers,


extensometers
excavation, replacement with lightweight
unload crest
fill
re-grade slope
create a new stable slope
Reduce driving forces
surface drainage
trench drains, interceptors, fill cracks
horizontal drains, drilled wells, tunnels,
subsurface drainage
adits
construct toe berm/ shear fill placement as buttress; increase shear
key
resistance below shear plane
surface treatment
rock fill blanket
Increase resisting
install anchors
drill and grout steel bars
install piles
drilled or driven piles
forces
MSE walls, gravity walls, pile walls,
retaining structure
tieback walls
Active methods
deep soil mixing
slurry mixtures
lower water table with drains; vacuum
drain subsurface
dewatering, siphoning
in-situ reinforcement
installation of soils nails or micropiles
Increase internal
reinforced backfill
geosynthetic reinforced soil
bioengineering
root support and drainage
strength
chemical stabilization
mix soils with additives
electro-osmosis
increase strength of clays
thermal treatment
artificially dry or freeze soil
deflect slide material around, above or
slide deflection
below structure
Protection measures
catchment basin or series of basins to
catchment basin
collect debris, reduce volume of slide.
instrumentation

6.2.1.2. Regulatory Methods


Regulatory methods may be used in local areas or regional districts that have developed
guidelines or rules for landslide risk assessment and stability analyses. For example, in
British Columbia, APEGBC has published a document entitled Guidelines for Legislated
Landslide Assessments for Proposed Residential Development in British Columbia. These
guidelines provide guidance for professionals carrying out stability analyses for proposed
37

residential developments and indicate how to relate the results of these analyses to a level of
landslide safety for residential development when required by provincial legislation.
In 2005, the Qubec government produced a guideline on the interpretation and use of
existing landslide maps in the Saguenay- Lac-Saint-Jean area (Qubec, 2005). The intent of
this guideline is to help local and regional authorities interpret and use the hazard maps
produced by the government in order to improve safety and control land-use. The guidelines
discussed risk management tools and cover common causes and triggers for the most common
types of landslides in the region. The guidelines also provide the minimum requirements that
will necessitate a geotechnical report when further investigation is required.
Jurisdictions can control land-use in landslide prone terrain. This can be accomplished by
refusing to allow development in areas of landslide susceptibility or by imposing covenants or
restrictions on the types of developments that are allowed. Mitigative measures may also be
required to allow the proposed development to proceed.
For example, the New Zealand Geological and Nuclear Sciences (GNS) published a
document entitled Guidelines for assessing planning policy and consent requirements for
landslide prone land (Saunders and Glassey, 2007). This document is primarily aimed at
educating land-use planners with the concepts and relevant issues to be considered when
incorporating landslide hazard information and assessment in the planning process. The
guidelines outline the criteria used to assess landslide hazards, provide examples of issues,
objectives, policies, rules, and assessment criteria. The guidelines discuss landslide concepts
to assist planners in understanding landslide processes in addition to proposing a risk-based
approach to land-use planning and approval.
Another example is the Land management handbook produced by the British Columbia
Ministry of Forests (Chatwin et al., 1994). The document is targeted to field personnel
primarily in the forestry industry who are operating in areas with existing or potential slope
stability problems. The handbook covers landslide processes and characteristics, techniques
for recognizing landslide-prone terrain, mitigation measures to manage unstable terrain, and
road deactivation and re-vegetation of unstable terrain.
A more general treatise on landslide studies for a North American audience primarily
targeting the non-specialist (home owners, developers, planners and others) that addresses all
landslide aspects from identification to monitoring and mitigation is the joint Geological
Survey of Canada and US Geological Survey treatise freely available online in several
languages (Highland and Bobrowsky, 2008; http://pubs.usgs.gov/circ/1325/).
6.2.1.3. Non-Regulatory Methods
Non-regulatory methods are used to encourage avoidance of the potential landslide risk. Local
jurisdictions can prevent or reduce potential development on landslide susceptible terrain by
purchasing the unstable terrain for parkland, swapping the land for an area of more suitable
terrain, or allowing increased density on safer portions of the land while imposing covenants
and/or restrictions on the more vulnerable areas. Governments can also provide financial
incentives to encourage developers to locate structures in safer areas or for environmental
protection of landslide prone terrain. Education and public awareness can also provide
incentive for developers to reduce the potential risk as it may be difficult to sell developments
in know hazardous areas.

38

6.2.2. Avoidance
In some cases, the risk of landslide occurrence is high and avoidance of development is the
best approach. Avoidance may involve relocation of the route or structure if the project is still
in the planning stages. In the cases where there are few location constraints to the project or as
noted above, if the project is still in the planning stages, avoidance may be highly cost
effective as the location with little or no risk can be selected. Alternatively, where
development has begun or constraints are in place on the resultant structure, avoidance may
not be feasible. It can also involve construction of a facility such as bridge or tunnel to avoid
the affected area. Avoidance may also involve the removal of unstable ground.
Avoiding unstable ground can be used effectively when geological reconnaissance or
investigation has delineated a landslide area. The earlier the landslide is identified, the easier
and less costly it will be to avoid. Avoidance can involve relocating a linear route or a specific
structure away from the affected area or it can involve bridging over the unstable zone.
6.2.3. Monitoring and Warning
Managing the risk through the use of slope monitoring or early warning systems is commonly
used on projects where there is some knowledge of the slope instability and it is not cost
effective to relocate the structure or linear feature. There are many techniques that can be used
for slope movement monitoring including conventional survey hubs, differential global
positioning systems (dGPS), LiDAR (land based and aerial), slope inclinometers, tiltmeters,
piezometers, extensometers, strain meters, acoustic instruments and vibration meters. All of
these systems can be used independently, or often in tandem, to monitor movements of an
existing or potential slide.
Warning systems can be set up to warn of impending failures. Real time systems that
monitor rainfall have been implemented in several jurisdictions (for example, Hong Kong and
San Francisco). Procedures are typically based on empirical and theoretical relations between
rainfall intensity and landslide activity. Such systems can trigger sirens, flashing lights or
radio and television broadcasts to warn local residents or operators of transportation systems.
The Alberta Geological Survey (AGS) has undertaken a long-term monitoring strategy of
the South Peak on Turtle Mountain that includes an early warning to residences of a possible
catastrophic rock avalanches. AGS has installed a variety of sensors including, crackmeters,
tiltmeters, extensometers, reflective prisms and dGPS receivers. The sensors are monitored in
real time and a warning management system automatically notifies officials of any change in
the mountain that warrants further investigation by experts.
6.3.

Active Methods

Active methods typically involve some form of construction to stabilize or reduce the
potential for a slope failure to occur. Methods of stabilization are typically divided into
reducing the driving forces, increasing the resisting forces or increasing the internal strength
of the slide mass.

39

6.3.1. Reduce the Driving Forces


Driving forces are those which tend to cause movement or drive the slope to failure. These
include the weight of the soil in the upper portion, the weight of surface water and vegetation
on the upper slope and porewater pressure acting on tension cracks. It can also include
flattening the slope and/or decreasing its height.
6.3.1.1. Unload Crest
The most common method of reducing the driving forces is to unload the slope by removing
material from the crest area. This also flattens the overall slope angle and typically increases
the slope stability. Slope stability sensitivity analyses can help determine the amount of
material that needs to be removed to achieve a desired factor of safety. If this material can be
placed at the base of the slope as a toe berm, this will increase the effectiveness of the
excavation. If the removed material has some potential use, it will increase the economic
viability of this option. In some cases, it may be required to replace the excavated material
with a lighter material, such as pumice stone, foamed concrete or Styrofoam (extruded
polystyrene). In 1997 in Summerland, BC unloading is the crest occurred at the North Beach
Rock Slide on Highway 97 near where about 50,000 cubic metres of rock was removed from
the crest of a large planar rock slide to reduce the movement to very small values.
6.3.1.2. Re-grade Slope
Slope re-grading usually consists of selective removal and filling of portions of the slope to
create a more stable configuration. Re-grading can result in a reduction of driving forces.
Early in the design phase, it may also be possible to raise the grade of the structure or linear
feature to improve overall slope stability. This method is particularly effective in deep-seated
landslides. This can also be used to remove material that contributes to slope instability, such
as colluvium and landslide debris.
6.3.1.3. Surface Drainage
Drainage of the upper slope using trench drains (French Drains) is a very cost effective way to
reduce the driving forces. Trench drains are typically installed by excavating trenches from 1
to 5 m deep at regular spacing either parallel or perpendicular to the movement direction. The
trenches are excavated and backfilled with free draining gravel or drainage pipes that allow
surface water to be transported away from the slope. Other surface drainage improvements
can be achieved by re-grading of small gullies and directing surface water away from the area
of instability. Water may be carried in trenches or in surface conduits such as half pipes or
closed pipes.
The sealing of existing cracks and fissures with low permeability material is also an
important part of improving surface drainage. Drainage blankets consist of free draining
granular material can also be used to shed water off the slope. Surface drainage treatment can
be particularly effective when combined with a subsurface drainage technique.

40

6.3.1.4. Subsurface Drainage


Subsurface drainage to reduce driving forces can consist of drilled vertical wells or subhorizontal drains. Vertical wells can be very effective when horizontal layers of impermeable
material (such as bentonite seams) do not allow water to percolate down to the groundwater
table. Perched water on these impermeable layers can significantly reduce slope stability. At
the north portal of the Edmonton SLRT (spell out) Project, large diameter drilled vertical
drains were effective in reducing perched water levels above bentonite seams within the
slope.
Sub-horizontal drains drilled on the slope can also be effective to lower the groundwater
table or drain water from a sub-horizontal permeable layer that overlies an aquitard. These
drains are typically drilled at a 2 to 5 degree angle above horizontal and are fitted with slotted
PVC pipe. The pipes can discharge freely or be connected to a larger diameter collector pipe.
A success story for the use of subsurface drainage is the Downie Slide, located upstream
of the Revelstoke Dam in southwestern British Columbia (Imrie et al., 1992). The risk of
failure on the Revelstoke Dam was a concern based on the Vaiont Slide failure that occurred
in Italy in 1963 which destroyed the town of Longarone in the valley below the Vaiont Dam
and caused some 2000 casualties. To improve the stability of the slide, a series of drainage
adits and drain holes were constructed within the slide. Since installation, the groundwater
levels in the slope have been reduced to less than pre-reservoir levels. In addition, a
significant reduction in the rate of down-slope movement has occurred within the past 25
years of monitoring.
6.3.2. Increase the Resisting Forces
Increasing the resisting forces improves the stability of the slope. These techniques involve
applying a resisting force near the toe of the slide, constructing some form of retaining
structure or installing reinforcing elements such as anchors or piles through the slide mass.
6.3.2.1. Construct Toe Berm/Shear Key
The most common and typically the most economic external force method is to construct a toe
berm at the base of the slope. Toe berms are typically constructed of soil available locally and
may contain drainage layers and/or geotextile separation. If the material can be excavated
from the crest of the slope this will double the effectiveness as it will both reduce the driving
force and increase the resisting force. The toe berm or buttress must be designed to be stable
against sliding along its base and against bearing failure of the additional weight. The buttress
must also be checked for failure within in its own mass. The size of the toe berm required is
estimated from the results of stability analyses. If there is inadequate room or soil support for
a toe berm, structural systems such as anchors, piles or retaining walls may be required. A
general guideline is that the width of the buttress typically ranges from about one third to
almost the full height of the slope being buttressed.
In addition, a shear key can be used to increase the shear resistance of the slope. In this
case, a key is excavated in to competent material below the failure surface. This key is then
backfilled with compacted fill, slurry, or concrete.

41

6.3.2.2. Surface Treatment


Surface treatments can be effective for small shallow type failures in soils that are susceptible
to erosion such as silts and fine sands. Such treatments usually involve placement of a blanket
of blasted rock fill or coarse granular fill on the surface of the sliding mass. This blanket
improves drainage, reduces the potential for surface erosion and increases the overall friction
angle along the potential failure plane. Where working room is limited, sometimes a layer of
the existing slope material is removed to make room for the surface blanket of more
competent material.
6.3.2.3. Install Anchors
Anchors are commonly used to stabilize slopes by applying resisting forces to a potentially
moving slide mass. They are usually installed by drilling and grouting steel bars into the
slope. Anchors are probably more common on steeper slopes composed of more competent
soils where there is a limited right of way. The soil must have adequate strength to develop
the pullout capacity of the anchors. Stability analysis or deformation analysis is required to
design the length, size and spacing of the anchor elements. Anchors are commonly used for
stabilizing steep slopes along transportation corridors in steep mountainous terrain. They can
also be used in conjunction with shotcrete or high strength steel mesh facing to create
structural support.
6.3.2.4. Install Piles
Piles are becoming a more common means of stabilizing a slide mass. They are typically
drilled caissons filled with concrete and reinforcing steel, however, they can also consist of
driven piles. The piles extend across the potential sliding plane and increase the shear
resistance. The embedment depth is based on slope stability analysis. Piles can also be used to
construct a retaining structure at the toe of the slope. Pile walls can be cantilever structures or
tied back with anchors. Various specialized techniques have recently been developed for
analysing pile stabilized landslides. Alberta transportation has successfully used pile walls to
remediate and stabilize landslides affecting Alberta highways (Abdelaziz et al., 2011).
6.3.2.5. Retaining Structures
Where there is insufficient space to allow construction of a toe berm, it is often necessary to
design some form of structural retention system to improve stability. Rigid walls are not able
to tolerate large deformations and therefore tend to attract large forces. Walls that are more
flexible are usually preferred. Some of the more common types of walls that are used include
mechanically stabilized earth (MSE) walls, masonry block gravity walls, concrete gravity or
cantilever walls, gabion gravity walls, sheet pile walls, pile walls (cantilever or tie back),
slurry walls, shotcrete and anchor walls and soil nailed walls. Each of these wall systems have
distinct advantages and disadvantages depending on the slope geometry, soil and groundwater
conditions, environmental factor, space availability and economic constraints. A detailed
review of each option is beyond the scope of this document.

42

Recently, the use of MSE retaining walls has become more common due to their
flexibility and reasonable cost. The backfill behind the facing elements is reinforced with steel
trips or layers of geosynthetics. The reinforcement develops stress from the soil either by
friction along the surface or passive resistance on elements with width normal to the
movement direction. Such walls are particularly favourable in seismically active areas
because of their flexibility and ability to distribute loads. Facing materials vary from wrapped
geosynthetics to concrete masonry blocks to precast concrete panels. A trend of allowing the
face to be vegetated for a greener look is also becoming popular.
6.3.2.6. Deep Soil Mixing
Deep soil mixing is a relatively new technique that has been used to stabilize unstable soils.
The technique involves the installation of a series of soil cement mixed barrettes in the failed
slope area. Replacement of the weaker soil with soil cement elements increases the shearing
resistance across the potential failure plane.
6.3.3. Increase Internal Strength
Increasing the internal strength involves improving the overall strength of the slide mass. This
can be accomplished in many different ways such as installing bulk reinforcement into the
slide mass or by treating the soil mass in some way. Some of the more common methods for
improving the internal strength of the slide mass are discussed below.
6.3.3.1. Drain Subsurface
Subsurface drainage can be used to improve overall internal strength of a slide mass. This
usually involves lowering the water table with some form of deep wells or sub-horizontal
drains. The drains can be fitted with vacuum devices to increase effectiveness. Siphoning
techniques can also be used.
6.3.3.2. In-situ Reinforcement
Installation of in-situ reinforcement such as soil nails or micropiles is an effective way to
stabilize some soil masses. Soil nailing involves the installation of closely spaced steel bars,
cables or tubes that forms a coherent reinforced soil mass. Soil nailing differs from anchor
tiebacks in that the nails are not tensioned and act as passive dowels.
Reticulated root piles can also be used to provide slope stabilization. Root piles are
typically small diameter micropiles or pin piles that are installed with a single steel rod. The
piles are installed at various angles to create a monolithic block of reinforced soil that
supports the sliding mass. The root like soil reinforced zone obtains its strength form the
three dimensional geometry of the micropiles.
6.3.3.3. Reinforced Backfill
It is possible to create steep slopes using reinforced backfill. Reinforcement types are
typically divided into extensible (geosynthetics) and non-extensible (steel). Extensible
43

reinforcement consisting of various types of geosynthetics (HDPE High Density


Polyethylene, PET polyethylene terephthalate synthetics and others) allows more
deflections and is commonly used for retaining walls and steeps slopes less than about 8 to 10
metres high. Non-extensible reinforcement consisting of steel strips or steel rods is typically
used for higher walls or slopes where less deflection can be tolerated such as at bridge
abutments.
Reinforced slopes using layers of geosynthetics to increase stability are becoming more
common. Such slopes often have a facing consisting of wrapped geosynthetic or some form of
welded wire mesh basket. The trend of obtaining a green facing by allowing vegetation to be
grown on the slope or on benches between facing elements is becoming very popular.
6.3.3.4. Bioengineering
Slope stabilization using bioengineering has been in use for some time. Vegetation planted on
a slope can increase the stability through root support reinforcement, by reducing the amount
of water on the slope (evapotranspiration) and by allowing more effective drainage of the
slope. Grasses and woody plants are most commonly used for slope bioengineering. The use
of vegetation for stabilization is best suited on slopes that have shallow failure mechanisms or
erosion related concerns and is strongly affected by the climate and soil conditions of the local
area. Vetiver grass, for example, is one of the most common species used in tropical regions
around the world. It is best to consult local experts if bioengineering is being considered for a
slope stabilization project.
In recent times, the combination of vegetation and structural elements has gained
popularity. The vegetation can provide support to the near surface soils and the structural
elements such as live cribwalls, soft gabions or vegetated concrete blocks provide the
structural support.
6.3.3.5. Chemical Stabilization
Chemical treatment of soils has been used to stabilize failed soil embankments. The concepts
are similar to those used to treat poor quality subgrades for highways and airport construction.
Typically, a lime slurry or grout is injected into the soil under pressure using some form of
perforated nozzles. The slurry mix can either naturally penetrate fissures or cracks in the
ground or be mixed into the soil using a higher pressure nozzle (jet grouting). Such techniques
are more applicable in flatter terrain where access is relatively easy. These techniques have
been used in the southeastern United States and in the UK.
6.3.3.6. Electro-osmosis
Electro-osmosis has been used to increase the in-situ shear strength of silts and clays.
Typically used to stabilize embankment foundations, electro-osmosis has also been used to
stabilize slopes although it is not common due to operational costs and uncertainties about its
effectiveness. In field tests on soft Leda Clay in eastern Canada, the undrained shear strength
was increased by about 50% using electro-osmosis.

44

6.3.3.7. Thermal Treatment


It is possible to increase soil strength using high temperatures that will remove moisture
from the soil and potentially fuse fine grained particles together. This technique has not been
used much due to the high cost of thermal treatment.
Ground freezing has been used more, particularly for the collection of samples and for
tunneling. However, freezing is typically used for temporary conditions because it is very
costly to keep the ground frozen for any length of time. Temporary freezing does not usually
provide a permanent strength increase of the soil. However, freezing has been successful for
longer term applications with the use of thermo-syphons. Some examples include: The
hospital and visitors centre in Inuvik, Northwest Territories (Holubec, 2008).
6.3.4. Protection Measures
Where it is not possible to avoid unstable ground it may be feasible to construct protection
structures. Such structures may consist of dikes, berms or walls that deflect slide material
away from vulnerable structures or catchment basins that halt or slow down slide debris.
These methods are discussed below.
6.3.4.1. Slide Deflection
Slide deflection berms have been used on numerous projects to deflect debris flows,
landslides and snow avalanches around small developments, bridge structures and other
sensitive structures. This typically involves construction of a lengthy berm made from native
materials along one side of the channel or construction of a new channel. Deflection dikes can
be effective near the apex of alluvial fans to deflect debris in a controlled manner along one
side or both sides of the fan. An example of such a deflection dike is located at Port Alice, BC
on Vancouver Island.
Another example of a slide deflection berm is the Mt. Stephen deflection dike located in
the Rocky Mountains near Field, BC. A large rock fill berm channelizes avalanches and
debris flows over a concrete cut and cover tunnel constructed by CP Rail in the 1980s. The
structures have been effective at reducing the impact on the railway. However, periodic
maintenance is required.
6.3.4.2. Catchment Basins
Where avoidance or deflection is not practical, it may be possible to slow down or catch
debris flows or landslides using catchment basins, energy dissipaters and/or check dams.
These structures usually involve excavation of large basins used to reduce the speed of slides.
Check dams and energy dissipaters consisting of concrete structures, mounds, walls, piles or
concrete fins are also used to impede slide progress. Many examples of large concrete check
dams and catchment basins are located on the Sea to Sky Highway 99 between Vancouver
and Squamish, BC. These structures require regular maintenance to clean out trapped debris.

45

6.4.

Emergency Management

Emergency management of recent or on-going landslides will depend on the consequences of


failure and the risk to persons and infrastructure. Once a landslide is identified, a preliminary
risk evaluation must be conducted to determine a reasonable course of action. This risk
evaluation will look at:
likelihood of a failure or re-occurrence of an event,
the consequence of failure in terms of injury or death to persons and economic
loss,
the need to undertake immediate landslide monitoring and/or investigation and
analysis,
the need to notify all potential stakeholders,
the need to warn or evacuate people and/or to shut down or close infrastructure,
potential methods of stabilization or slope treatment.
The perception of the potential loss can also play an important role in the evaluation of a
landslide. Communication to all stakeholders (persons of interest) is critical in managing an
emergency.
Risk scenarios must also be evaluated. A landslide may trigger another event which could
lead to a catastrophic loss or damage to property. For example, a landslide could enter a large
body of water generating waves that could damage properties and structures a considerable
distance from the initial event. Risks must be examined in terms of costs and benefits to
determine if they are acceptable or if remedial action is needed. Once the risk has been
evaluated, it must be controlled to reduce the frequency and severity. Monitoring and warning
systems are often deployed as an interim measure until the risk evaluation and risk control
measures can be determined and implemented.
The final phase involves implementation of the risk control or action plan measures. This
can involve any of the treatment options or combination of options discussed in this
contribution. The components involved in slope stability emergency management are quite
extensive and not discussed in detail here. For a better review of this aspect of landslide study
refer to several contributions in Bobrowsky (2013).
A recent example of emergency management involved the development of a large rock
slide on Highway 97 in south-central British Columbia in 2008. The potential slide forced the
closure of the highway for 17 days while detailed monitoring and investigation programs were
carried out. There was potential for the slide to affect some houses below and for the slide to
enter a large lake causing wave damage. A detailed communication plan was developed
through the local media and nearby residents to ensure that all stakeholders were informed of
the process and timing for the action plan. The slide area was unloaded to reduce the
probability of catastrophic occurrence and to allow the highway to re-open.
6.5.

Limitations or Treatment Methods

All of the treatment options discussed in this section have limitations. Some of the more
important limitations are as follows:
knowing the size, extent and depth of the potential failure,
having adequate space to implement a specific remedial option,
having adequate resources to implement a specific remedial option,
46

choosing a method that adequately addresses the stability concerns,


choosing a method that keeps deformations to acceptable levels,
requiring significant monitoring or inspection,
having too high a cost to permit stabilization.
For most landslides, it is unlikely that all of the desired information will be available to
allow assessment and develop the best treatment option(s). Local experience is a key factor in
estimating some geological and environmental aspects of the slide conditions. Engineering
judgement is always required in landslide assessment, risk evaluation and treatment
application.

7.0

ACKNOWLEDGEMENTS

The authors appreciate and acknowledge the input from a great number of colleagues and
professionals involved in landslide studies. In particular we recognize the comments and
written contributions of Marc Andre Brideau and Scott McDougall for Section 4, Matthieu
Sturzenegger (LiDAR and photogrammetry), Vern Singhroy (InSAR), Matthias Jakob (data
connectivity), Francisco Moreno (early warning systems), and Doug Baker (in situ
techniques). Without their contribution and co-operation these sections would not have been
possible. Finally, we thank our individual employers BGC Engineering Inc., Government of
Canada, Thurber Engineering Ltd., Simon Fraser University and VanDine Geological
Engineering Ltd. for their support allowing us to work on this endeavor. This contribution
benefited from the review of David Huntley (GSC Vancouver).

8.0

REFERENCES

Abdelaziz, T., Proudfoot, D., and Skirrow, R., 2011. Stabilization of Alberta Highway
landslides using Pile Walls. in Proceedings of the 14th Pan-American Conference on Soil
Mechanics and Geotechnical Engineering and the 64th Canadian Geotechnical
Conference. Toronto, Ontario, Canada, 8 pages.
Abelln, A., Oppikofer, T., Jaboyedoff, M., Rosser, N.J., Lim, M., and Lato, M.J., 2014.
Terrestrial laser scanning of rock slope instabilities; Earth Surface Processes and
Landforms, v. 39, no. 1, p. 80-97.
Alzoubi, A.M., 2009. The effect of tensile strength on the stability of rock slope. Ph.D.
thesis, University of Alberta, Edmonton, Alberta, Canada.
Ambrosi, C. and Crosta, G.B., 2011. Valley shape influence on deformation mechanisms of
rock slopes. in Slope Tectonics, (ed. M. Jaboyedoff), Geological Society of London,
Special Publication 351, p. 215 233.
APEGBC, 2010. Guidelines for Legislated Landslide Assessments for Proposed residential
Developments in BC, Association of Professional Engineers and Geoscientists of British
Columbia, Revised May 2010, 75 pages.
ASTM, 2005. Standard test method for mechanical cone penetration tests of soil, ASTM
International, D3441-05.
ASTM, 2007. Standard test method for electronic friction cone and piezocone penetration
testing of soils, ASTM International, D5778-07.
ASTM, 2008a. Standard test method for soil compaction determination at shallow depths
using 5-lb (2.3 kg) dynamic cone penetrometer. ASTM International, D7380-08.
47

ASTM, 2008b. Standard test method for field vane shear test in cohesive soil, ASTM
International D2573-08.
ASTM, 2011. Standard test method for standard penetration test (SPT) and split-barrel
sampling of soils, ASTM International, D1586-11.
Australian Geomechanics Society (AGS), 2007. A National Landslide Risk Management
Framework for Australia; Australian Geomechanics, v. 42, p. 63-114
Alejano, L.R., Ferrero, A.M., Ramirez-Oyanguren, P., and Alvarez Fernandez, M.I., 2011.
Comparison of limit-equilibrium, numerical and physical models of wall slope stability;
International Journal of Rock Mechanics and Mining Sciences; v. 48, p. 16-26.
Aylsworth, J.M. and Hunter, J.A., R., 2003. The Ottawa Valley Landslide Project: A
geophysical and geotechnical investigation of geological controls on landsliding and
deformation in Leda Clay. in Proceedings, GeoHazards 2003, 3rd Canadian Conference
on Geohazards, Edmonton, Alberta, Canada, p. 227-234.
Ayotte, D. and Hungr, O., 2000. Calibration of a runout prediction model for debris-flows and
avalanches; in Proceedings of the Second International Conference on Debris-Flow
Hazards Mitigation, ed. G.F. Wieczorek and N.D. Naeser. A.A. Balkema, pp. 505-514.
Baker, D.G., 1991. Wahleach power tunnel monitoring; in Field Measurements in
Geotechnics, A.A. Balkema, Rotterdam, p. 467-479.
Baligh, M.M. and Azzouz, A.S., 1975. End effects of stability of cohesive slopes; Journal of
the Geotechnical Engineering Division v. 101, p. 1105-1117.
Bateson, L., Cigna, F., Boon, D., and Sowter, A., 2015. The application of the Intermittent
SBAS (ISBAS) InSAR method to the South Wales Coalfield, UK; International Journal
of Applied Earth Observation and Geoinformation, v. 34, p. 249-257.
Baum, R.L. and Godt, J.W., 2010. Early warning of rainfall-induced shallow landslides and
debris flows in the USA; Landslides, v. 7, p. 259-272.
BC MOF (BC Ministry of Forests), 2002. Forest road engineering guidebook, Forest Practices
Branch, Victoria, BC, 208 pages.
BGC Engineering Inc., 2005. Berkley landslide risk management (Project number 0404-006).
Report to the District of North Vancouver, 75 pages.
BGC Engineering Inc., 2006. Berkley Landslide Risk Management, Phase 1 Risk
Assessment; report to District of North Vancouver, January 13, 2006, 30 pages plus
appendices.
Bhasin, R. and Kaynia, A.M., 2004. Static and dynamic simulation of a 700-m-high rock
slope in western Norway; Engineering Geology, v. 71, p. 213 226.
Bichler, A., Bobrowsky, P., Best, M., Douma, M., Hunter, J., Calvert, T., and Burns, R.,
2004. Three-dimensional mapping of a landslide using a multi-geophysical approach: the
Quesnel Forks landslide; Landslides, v. 1, p. 29-40.
Bishop, A.W., 1955. The use of the slip circle in the stability analysis of slopes; Geotechnique
v. 5, p. 7-17.
Bobrowsky, P. (Editor) 2013. Encyclopedia of Natural Hazards. Springer, New York, 1141
pages.
Bobrowsky, P. and Couture, R. (2014) Landslide Terminology Canadian Technical
Guidelines and Best Practices related to Landslides: a national initiative for loss
reduction. Geological Survey of Canada, Open File 7623, 68 pages.

48

Bourdeau, C., Havenith, H-B., Fleurisson, J.E., and Grandjean, G., 2004. Numerical
modelling of seismic slope stability. in Engineering Geology for Infrastructure Planning
in Europe, a European Perspective, (ed.) R. Hack, R. Assam and R. Charlier. Lecture
Notes in Earth Sciences, v. 104, p. 671-684.
Bozzano, F., Lenti, L., Martino, S., Montagna, A., and Paciello, A., 2011. Earthquake
triggering of landslides in highly jointed rock masses: Reconstruction of the 1783 Scilla
rock avalanche (Italy); Geomorphology, v. 129, p. 294308.
Brideau, M.-A., 2010. Three-dimensional kinematic controls on rock slope stability
conditions. Ph.D. thesis. Simon Fraser University, Burnaby, British Columbia, Canada.
Calvetti, F., Crosta, G., and Tartarella, M., 2000. Numerical simulation of dry granular flows:
from the reproduction of small-scale experiments to the prediction of rock avalanches;
Rivista Italiana di Geotecnico, v. 34, p. 21-38.
Cannon, S.H., 1993. An empirical model for the volume-change behavior of debris flows; in
Proceedings of Hydraulic Engineering 93, (ed.) H.W. Shen, S.T. Su and F. Wen.
American Society of Civil Engineers, v. 2, p. 1768-1773.
Cappa, F., Guglielmi, Y., Soukatchoff, V.M., Mudry, J., Bertrand, C., and Charmoille, A.,
2004. Hydromechanical modeling of a large moving rock slope inferred from slope
levelling coupled to spring long-term hydrochemical monitoring: Example of the La
Clapire landslide (Southern Alps, France); Journal of Hydrology, v. 291, p. 67 90.
Cavers, D.S., 1981. Simple methods to analyze buckling of rocks slopes; Rock Mechanics, v.
14, p. 87-104.
Cavounidis, S., 1987. On the ratio of factor of safety in slope stability analyses;
Geotechnique, v. 37, p. 207-210.
Cepeda, J., Chvez, J.A., and Cruz Martnez, C., 2010. Procedure for the selection of runout
model parameters from landslide back-analyses: application to the metropolitan area of
San Salvador, El Salvador; Landslides, v. 7, p. 105-116.
Chatwin, S.C., Howes, D.E., Schwab, J.W., and Swanston, D.N., 1994. A guide for
management of landslide-prone terrain in the Pacific Northwest; BC Ministry of Forests,
Land Management Handbook 18, 220 pages.
Chemenda, A.I., Bois, T., Bouissou, S., and Tric, E., 2009. Numerical modelling of the
gravity-induced destabilization of a slope: The example of the La Clapire landslide,
southern France; Geomorphology, v. 109, p. 86 93.
Chen, R.H. and Chameau, J.-L., 1982. Three-dimensional limit equilibrium analysis of slopes;
Geotechnique, v. 32, p. 31-40.
Cornforth, D.H., 2005. Landslides in Practice: Investigation, Analysis, and
Remedial/Preventative Options in Soils, Wiley, 596 pages.
Corominas, J., 1996. The angle of reach as a mobility index for small and large landslides;
Canadian Geotechnical Journal, v. 33, p. 260-271.
Crawford, C.B. and Eden, W.J., 1967. Stability of natural slopes in sensitive clay. Journal of
Soil Mechanics and Foundations Division, v. 93, no. SM4, p. 419-436.
Cundall, P.A. and Damjanac, B., 2009. A comprehensive 3D model for rock slopes based on
micromechanics. in Proceedings of the Slope Stability 2009 Conference, Santiago, Chile,
10 pages.
Cundall, P.A. and Strack, O.D.L., 1979. A discrete numerical model for granular assemblies;
Geotechnique, v. 29, p. 47-65.

49

Dalbey, K., Patra, A.K., Pitman, E.B., Bursik, M.I., and Sheridan, M.F., 2008. Input
uncertainty propagation methods and hazard mapping of geophysical mass flows; Journal
of Geophysical Research, v. 113, no. B05203, 16 pages.
Dawson, E.M., Roth, W.H., and Drescher, A., 1999. Slope stability analysis by strength
reduction; Geotechnique, v. 49, p. 835 840.
Delacourt, C., Allemand, P., Berthier, E., Raucoules, D., Casson, B., Grandjean, P., Pambrun,
C., and Varel, E., 2007. Remote-sensing techniques for analysing landslide kinematics: a
review; Bulletin De La Societe Geologique De France, v. 178, p. 89-100.
Diederichs, M.S., Lato, M., Hammah, R., and Quinn, P. 2007. Shear strength reduction
approach for slope stability analyses. in Proceedings of the First CanadianUS Rock
Mechanics Symposium, (ed.) E. Eberhardt, D. Stead and T. Morrison. A. A. Balkema,
Rotterdam, p. 319327.
Discenza, M.E., Esposito, C., Martino, S., Petitt, M., Prestininzi, A., and Scarascia Mugnozza,
G., 2011. The gravitational slope deformation of Mt. Rocchetta ridge (central Apennines,
Italy): geological-evolutionary model and numerical analysis; Bulletin Engineering
Geology and the Environment, v. 70, p. 559575.
Duncan, J.M., 1996. Soil Slope Stabiltiy Analysis; in Landslides Invenstigation and
Mitigation, (ed.) A.K. Turner and R.L. Schuster. Transportation Research Board Special
Report 247, p. 337-371.
Dunn, D. and Ricketts, B., 1994. Surficial geology of Fraser Lowlands digitized from GSC
Maps 1484A, 1485A, 1486A, and 1487A (92 G/1,2,3,6,7; 92 H/4). Geological Survey of
Canada, Open File 2894.
Dunnicliff, J., 1993. Geotechnical Instrumentation for Monitoring Field Performance. John
Wiley & Sons, New York. 577 pages.
Eberhardt, E. and Stead, D., 1998. Numerical analysis of slope instability in thinly bedded
weak rock. in Proceedings of the 8th International Association of Engineering Geology
Congress, Vancouver, British Columbia, Canada, p. 30113018.
Eberhardt, E., Stead, D., and Coggan, J.S., 2004. Numerical analysis of initiation and
progressive failure in natural rock slopes: The 1991 Randa rockslide; International Journal
of Rock Mechanics and Mining Science, v. 41, p. 69 87.
Eberhardt, E., Watson, A.D., and Loew, S., 2008. Improving the interpretation of slope
monitoring and early warning data through better understanding of complex deep-seated
landslide failure mechanisms; in Landslides and Engineered Slopes: From the Past to the
Future, 10th International Symposium on Landslides and Engineered Slopes, (eds.) Z.
Chen, J. Zhang, Z. Li, F. Wu, and K. Ho. Taylor & Francis, London, p. 39-51.
Environment Yukon, 2012. 30 meter Yukon Digital Elevation Model.
www.env.gov.yk.ca/mapspublications/geomatics/data/30m_dem.php#50kntdb. Updated
20 September 2012. Accessed 22 September 2012.
Evans, S.G. and Hungr, O., 1993. The assessment of rockfall hazard at the base of talus
slopes; Canadian Geotechnical Journal, v. 30, p. 620-636.
Fannin, R.J. and Wise, M.P., 2001. An empirical-statistical model for debris flow travel
distance; Canadian Geotechnical Journal, v. 38, p. 982-994.
Fell, R., Ho, K.K.S., Lacasse, S., and Leroi, E., 2005. A framework for landslide risk
assessment and management; in Proceedings of the International Conference on
Landslide Risk Management, Vancouver, Canada, (eds.) O. Hungr, R. Fell, R. Couture,
and E. Eberhardt. A.A. Balkema, Netherlands, p. 3-25.
50

FLO-2D Software Inc., 2007. FLO-2D Users Manual Version 2007.06; www.flo-2d.com.
Fredlund, D.G. and Krahn, J. 1977. Comparison of slope stability methods of analysis;
Canadian Geotechnical Journal, v. 14, p. 429-439.
Friele, P.A. and Clague, J.J., 2005. Multifaceted hazard assessment of Cheekeye fan, a large
debris-flow fan in south-western, British Columbia; in Debris-flow Hazards and Related
Phenomena, (eds.) M. Jakob and O. Hungr. Praxis/Springer, p. 659-683.
Froese, C.R. and Moreno, F., 2014. Structure and components for the emergency response
and warning system on Turtle Mountain; Natural Hazards, v. 70, no. 3, p. 1689-1712.
Froese, C.R., Charriere, M., Humair, F., Jaboyedoff, M., and Pedrazzini, A., 2012.
Characterization and management of rockslide hazard at Turtle Mountain, Alberta,
Canada. in Landslides: Types, Mechanisms and Modeling, (eds.) J.J. Clague and D.
Stead. Cambridge University Press, Chicago, p. 282-296.
Galas, S., Dalbey, K., Kumar, D., Patra, A., and Sheridan, M., 2007. Benchmarking
TITAN2D mass flow model against a sand flow experiment and the 1903 Frank Slide; in
Proceedings of the 2007 International Forum on Landslide Disaster Management, (ed.) K.
Ho and Z. Li. Hong Kong Geotechnical Engineering Office, p. 899-918.
Geertsema, M., and Torrance, J.K., 2005. Quick Clay from the Mink Creek Landslide near
Terrace, British Columbia: Geotechnical Properties, Mineralogy, and Geochemistry;
Canadian Geotechnical Journal, v. 42, p. 907-918.
Gerath, R., Jakob, M., Mitchell, P., VanDine, D., Finn, L., Gillespie, D., Kuan, S. Naesgaard,
E., Patrick, B., Skermer, N., and Wallis, D., 2010. Guidelines for legislated landslide
assessments for proposed residential development in British Columbia. Association of
Professional Engineers and Geoscientists of British Columbia, 76 pages.
Gischig, V.S., 2011. Kinematics and failure mechanisms of the Randa rock slope instability
(Switzerland) Ph.D. thesis, ETH Zurich, Switzerland, no. 19730, 204 pages.
Golder Associates, 2012. Fracman 7.40 http://fracman.golder.com/
Gonzlez, E., Herreros, M.I., Pastor, M., Quecedo, M., and Fernndez Merodo, J.A., 2003.
Discrete and continuum approaches for fast landslide modelling; in Proceedings of the 1st
International PFC Symposium, (ed.) H. Konietsky. Swets and Zeitlinger, Lisse, p. 307313.
Gonzalez de Vallego, L.I. and Ferrer, M., 2011. Geological Engineering, CRC Press, 678
pages.
Goodman, R.E. and Bray, J.W., 1976. Toppling of rock slopes; in Specialty Conference on
Rock Engineering for Foundations and Slopes. American Society of Civil Engineering,
Boulder, Colorado, p. 201-234.
Grenon, M. and Hadjigeorgiou, J., 2010. Integrated structural stability analysis for
preliminary open pit design; International Journal of Rock Mechanics and Mining
Sciences v. 47, p. 450-460.
Griffiths, D.V. and Lane, P.A., 1999. Slope stability analysis by finite elements;
Geotechnique, v. 49, p. 387-403.
Griffiths, D.V. and Marquez, R.M., 2007. Three-dimensional slope stability analysis by
elasto-plastic finite elements; Geotechnique, v. 57, p. 537-546.
Griswold, J.P., 2004. Mobility Statistics and Hazard Mapping for Non-volcanic Debris Flows
and Rock Avalanches. M.Sc. thesis, Portland State University, Portland, Oregon, USA.

51

Grneng, G., Lu, M., Nilsen, B., and Jenssen, A.K., 2010. Modelling of time-dependent
behavior of the basal sliding surface of the knes rockslide area in western Norway.
Engineering Geology, v. 114, p. 442-422.
Guglielmi, Y., Capp, F.J., Rutqvist, J., Tsang, C.-F., and Thoraval, A., 2008. Mesoscale
characterization of coupled hydromechanical behavior of a fractured-porous slope in
response to free water surface movement; International Journal of Rock Mechanics and
Mining Science, v. 45, p. 862-878.
Guzzetti F., Stark C., and Salvati P., 2005. Evaluation of Flood and Landslide Risk to the
Population of Italy; Environmental Management, v. 1, p. 15-36.
Guzzetti, F., Crosta, G., Detti, R., and Agliardi, F., 2002. STONE: a computer program for the
three-dimensional simulation of rock falls; Computers and Geosciences, v. 28, p. 10791093.
Hammah, R.E., Yacoub, T., and Curran, J.H., 2009a. Variation of failure mechanisms of
slopes in jointed rock masses with changing scale, ROCKENG09. in Proceedings of the
3rd CANUS Rock Mechanics Symposium, (ed.) M. Diederichs and G. Grasselli. Toronto,
Ontario, Canada. Canadian Association of Rock Mechanics, Paper 3956.
Hammah, R.E., Yacoub, T. and Curran, J.H., 2009b. Probabilistic slope analysis with the
finite element method. in U.S. Rock Mechanics Symposium and 4th U.S.Canada Rock
Mechanics Symposium, ARMA 09149.
Hanssen, R.F., 2001. Radar Interferometry: Data Interpretation and Error Analysis. Kluwer
Academic Publishers, Dordrecht, 308 pages.
Hawley, P.M., Martin, D.C., and Acott, C.P., 1986. Failure mechanics and design
considerations for footwall slopes; CIM Bulletin, v. 79, p. 47-53.
Havenith, H.-B., Jongmans, D., Faccioli, E., Abdrakhmatov, K., and Bard, P.Y., 2002. Site
effect analysis around the seismically induced Ananevo rockslide, Kyrgyzstan, Bulletin
Seismological Society of America, v. 92, p. 3190-3209.
Hensold, G., 2011. An integrated study of deep-seated gravitational slope deformations at
Handcar Peak, Southwestern British Columbia. M.Sc. thesis, Simon Fraser University,
British Columbia, Canada, 135 pages.
Highland, L.M. and Bobrowsky, P., 2008. The Landslide Handbook: a guide to understanding
landslides. U.S. Geological Survey, Circular 1325, 129 pages.
Holobec, I., 2008. Flat Loop Thermosyphon Foundations in Warm Permafrost. Report
prepared for Government of the Northwest Territories, 119 pages.
Hsu, K.J., 1975. Catastrophic debris streams (sturzstroms) generated by rockfalls; Geological
Society of America Bulletin, v. 86, p. 129-140.
Hungr, O., 1981. Dynamics of rock avalanches and other types of slope movements; Ph.D.
thesis, University of Alberta, Edmonton, Alberta, Canada, 506 pages.
Hungr, O., 1987. An extension of Bishop's simplified method of slope stability analysis to
three dimensions; Geotechnique, v. 37, p. 113-117.
Hungr, O., 1990. Mobility of rock avalanches; Report of the National Research Institute for
Earth Science and Disaster Prevention, Japan, v. 46, p. 11-20.
Hungr, O., 1994. A general limit equilibrium model for three-dimensional slope stability
analysis: Discussion; Canadian Geotechnical Journal, v. 31, p. 793-795.
Hungr, O., 1995. A model for the runout analysis of rapid flow slides, debris flows, and
avalanches; Canadian Geotechnical Journal, v. 32, p. 610-623.

52

Hungr, O., 1997. Slope stability analysis; in 2nd Pan-American Symposium on Landslides,
Rio de Janeiro, Brazil, p. 123-136.
Hungr, O. and Amann, F., 2011. Limit equilibrium of asymmetric laterally constrained
rockslides; International Journal of Rock Mechanics and Mining Sciences, v. 48, p. 748758.
Hungr, O., Morgan, G.C., and Kellerhals, R., 1984. Quantitative analysis of debris torrent
hazards for design of remedial measures; Canadian Geotechnical Journal, v. 21, p. 663677.
Hungr, O. and Evans, S.G., 1996. Rock avalanche runout prediction using a dynamic model;
in Proceedings of the 7th International Symposium on Landslides, ed. K. Senneset. A.A.
Balkema, p. 233-238.
Hungr, O. and McClung, D.M., 1987. An equation for calculating snow avalanche run-up
against barriers. In Proceedings of the Symposium on Avalanche Formation, Movement
and Effects, Davos. International Association of Hydrological Sciences, Publication 162,
pp. 605-612.
Hungr, O., Morgenstern, N., and Wong, H.N., 2007. Review of benchmarking exercises on
landslide debris runout and mobility modelling; in Proceedings of the 2007 International
Forum on Landslide Disaster Management; ed. Ho and Li. Hong Kong Geotechnical
Engineering Office, p. 755-812.
Hungr, O., Salgado, F.M., and Byrne, P.M., 1989. Evaluation of a three-dimensional method
of slope stability analysis; Canadian Geotechnical Journal, v. 26, p. 679-686.
Hunt, R.E., 2007. Geotechnical Investigation Methods: A Field Guide for Geotechnical
Engineers, CRC Press.
Hunter, G. and Fell, R. 2003. Travel distance angle for rapid landslides in constructed and
natural soil slopes; Canadian Geotechnical Journal, v. 40, p. 1123-1141.
ICOLD (International Committee on Large Dams, 1989. Monitoring of dams and foundations;
ICOLD Bulletin 68.
Imrie, A.S., Moore, D.P., and Enegren, E.G., 1992. Performance and maintenance of the
drainage system at Downie Slide; in Proceedings of the 6th International Symposium on
Landslides, Christchurch, New Zealand, p. 751-757.
Itasca, 2011. FLAC 7.0, FLAC3D V4, UDEC 5.0, 3DEC 4.0, PFC 4.0, PFC3D 4.0.
Minneapolis, MN: Itasca Consulting Group Inc . [www.itascacg.com].
Iverson, R.M. 1997. The physics of debris flows. Review of Geophysics, v. 35, no. 3, p. 245296.
Iverson, R.M., Schilling, S.P. and Vallance, J.W. 1998. Objective delineation of lahar-hazard
inundation zones; Geological Survey of America Bulletin, v. 110, no. 8, p. 972-984.
Jaboyedoff, M., Derron, M-H., Jakubowski, J., Oppikofer, T., and Pedrazzini, A., 2012. The
2006 Eiger rock slide, European Alps. in Landslides: Types, Mechanisms and Modeling,
(ed.) J.J. Clague and D. Stead. Cambridge University Press, p. 282-296.
Jackson, L.E. Jr., Bobrowsky, P.T. and Bichler, A., 2012. Identification, Maps and Mapping
Canadian Technical Guidelines and Best Practices related to Landslides: a national
initiative for loss reduction. Geological Survey of Canada, Open File 7059, 33 pages.
Jakob M., Owen T., and Simpson T., 2011. A regional real-time debris-flow warning system
for the District of North Vancouver, Canada; Landslides, v. 9, no. 2, p. 165-178.
Janbu, N., 1973. Slope stability computation; in Embankment Dam Engineering, Casagrande
Volume, (ed.) R.C. Hirschfeld and S.J. Poulos. John Wiley & Sons, New York, p. 47-86.
53

Jefferies, M., Lorig, L., and Alvarez, C., 2008. Influence of rock strength spatial variability on
slope stability. in Continuum and Distinct Element Numerical Modeling in GeoEngineering 2008, (ed.) R. Hart, C. Detournay and P. Cundall. Minneapolis, Minnesota,
USA, Itasca Consulting Group, Paper 0105.
Jibson, R.W., 2012. Models of triggering landslides during earthquakes. in Landslides: Types,
Mechanisms and Modeling, (ed.) J.J. Clague and D. Stead. Cambridge University Press,
p. 196-206.
Jibson, R.W. and Jibson, M.W., 2003. Java Programs for using Newmarks Method and
Simplified Decoupled Analysis to Model Slope Performance during Earthquakes, USGS.
Open File Report 03-005.
Jing, L. and Husdon, J.H., 2002. Numerical methods in rock mechanics. International Journal
of Rock Mechanics and Mining sciences, v. 39, no. 4, p. 409-27.
Jing, L. and Stephansson, O., 2007. Fundamentals of Discrete Element Methods for Rock
Engineering: Theory and Applications; Developments in Geotechnical Engineering, v.
85, p. 1-21.
Johnson, R.B. and DeGraff, J.V., 1988. Principles of engineering geology. J. Wiley & Sons,
New York, 497 pages.
Kalenchuk, K.S., 2010. Multi-dimensional analysis of large, complex slope instability. Ph.D.
thesis, Queens University, Kingston, Ontario, Canada.
Krahn, J. 2003. The limits of limit equilibrium analyses; Canadian Geotechnical Journal, v.
40, p. 643-660.
Khan, M.S., 2010. Investigation of discontinuous deformation analysis for application in
jointed rock masses. Ph.D. thesis, University of Toronto, Toronto, Ontario, Canada.
Kromer, R.A., Hutchinson, D.J., Lato, M.J., Gauthier, D., and Edwards, T., 2015. Identifying
Rock Slope Failure Precursors Using LiDAR for Transportation Corridor Hazard
Management. Engineering Geology, v. 195(C), p. 93-103.
Kveldsvik, V., 2008. Static and dynamic stability analyses of the 800m high knes rock
slope, western Norway. PhD thesis, Department of Geology and Mineral Resources
Engineering, Norwegian University of Science and Technology, Trondheim
Lato, M.J., Hutchinson, D.J., Gauthier, D., Edwards, T., and Ondercin, M., 2014. Comparison
of airborne laser scanning, terrestrial laser scanning, and terrestrial photogrammetry for
mapping differential slope change in mountainous terrain; Canadian Geotechnical
Journal, v. 52, p. 1-12.
Lato M., Diederichs M.S., Hutchinson D.J., and Harrap R., 2011. Evaluating roadside rock
masses for rockfall hazards using LiDAR data: optimizing data collection and processing
protocols; Natural Hazards, v. 60, p. 831-864.
Lato, M.J., Gauthier, D., and Hutchinson, D.J. 2015. Selecting the Optimal 3D Remote
Sensing Technology for the Mapping, Monitoring and Management of Steep Rock Slopes
Along Transportation Corridors. in Transportation Research Board 94th Annual
Meeting (No. 15-3055).
Legros, F., 2002. The mobility of long-runout landslides; Engineering Geology, v. 63, p. 301331.
Leith, K., 2012. Stress development and geomechanical controls on the geomorphic evolution
of alpine valleys, Ph.D. thesis, ETH Zurich, Switzerland, 169 pages.
Li, T., 1983. A mathematical model for predicting the extent of a major rockfall; Zeitschrift
fr Geomorphologie, v. 27, p. 473-482.
54

Lillesand, T.M., Kiefer, R.W., and Chipman, J.W., 2008. Remote sensing and image
interpretation. Sixth ed. John Wiley & Sons, New York, 756 pages.
Mahmoud, M., Woeller, D., and Robertson, P.K., 2000. Detection of shear zones in a natural
clay slope using the cone penetration test and continuous dynamic sampling; Canadian
Geotechnical Journal, v. 37, p. 652-661.
Massonnet, D. and Feigl, K., 1998. Radar interferometry and its applications to changes in the
Earths surface; Reviews in Geophysics, v. 36, p. 441-500.
Mears, A.I., 1981. Design criteria for avalanche control structures in the runout zone. United
States Department of Agriculture, Forest Service, Rocky Mountain Forest and Range
Experimental Station, General Technical Report RM-84, 28 pages.
McClung, D.M. and Mears, A.I., 1995. Dry-flowing avalanche run-up and runout; Journal of
Glaciology, v. 41, p. 359-372.
McDougall, S., 2006. A New Continuum Dynamic Model for the Analysis of Extremely
Rapid Landslide Motion across Complex 3-D Terrain; Ph.D. thesis, University of British
Columbia, Vancouver, British Columbia, Canada.
McDougall, S., McKinnon, M., and Hungr, O., 2012. Developments in landslide runout
prediction. in Landslides: types, mechanisms and modeling, (ed.) J.J. Clague and D.
Stead. Cambridge University Press, Cambridge, p. 187-195.
McGuffey, V.C., Modeer, V.A, and Turner, A.K., 1996. Subsurface exploration. in
Landslides: Investigation and Mitigation, (ed.) A.K. Turner and R.L. Schuster.
Transportation Research Board, Special Report 247, p. 231-277.
McKinnon, M., 2010. Landslide Runout: Statistical Analysis of Physical Characteristics and
Model Parameters; M.Sc. thesis, University of British Columbia, Vancouver, British
Columbia, Canada.
Moore, D.P., Imrie, A.S., and Baker, D.G., 1991. Rockslide risk reduction using monitoring;
in Proceedings of Canadian Dam Safety Conference, Whistler BC, Canadian Dam Safety
Association.
Moore, J.R., Gischig, V., Burjanek, J., Loew, S., and Fh, D., 2011. Site Effects in Unstable
Rock Slopes: Dynamic Behavior of the Randa Instability (Switzerland), Short Note;
Bulletin of the Seismological Society of America, v. 101, no. 6, p. 31103116.
Moore, J.R., Gischig, V., Amann, F., Hunziker, M., and Burjanek, J., 2012. Earthquaketriggered rock slope failures: Damage and site effects. in Proceedings, 11th International
and 2nd North American Symposium on Landslides and Engineered Slopes; Protecting
Society through Improved Understanding, Banff, Canada, v. 1, p. 869-74.
Morgenstern, N.R. and Price, V.E., 1965. The analysis of the stability of general slip surfaces;
Geotechnique, v. 15, p. 79-93.
Mosher, D.C., Monahan, P.A., Barrie, J.V., and Courtney, R.C., 2004. Coastal submarine
failures in the Strait of Georgia, British Columbia: Landslide of the 1946 Vancouver
Island Earthquake; Journal of Coastal Research, v. 20, no. 1, p. 277-291.
Naesgaard, E., 2011. A hybrid effective stress total stress procedure for analyzing soil
embankments subjected to potential liquefaction and flow, PhD. Thesis, Department of
Civil Engineering, University of British Columbia, Canada, 192 pages.
Nasmith, H. W. and Mercer, A.G., 1979. Design of dykes to protect against debris flows at
Port Alice, British Columbia; Canadian Geotechnical Journal, v. 16, no. 4, p. 748-757.
Newmark, N.M., 1965. Effects of earthquakes on dams and embankments, Geotechnique,
15,139-159
55

Nicoletti, P.G. and Sorriso-Valvo, M., 1991. Geomorphic controls of the shape and mobility
of rock avalanches; Geological Society of America Bulletin, v. 103, p. 1365-1373.
Pedrazzini, A., Abelln, A., Jaboyedoff, M., and Oppikofer, T., 2011. Monitoring and failure
mechanism interpretation of an unstable slope in Southern Switzerland based on
terrestrial laser scanner; in Proceedings of the 64th Canadian Geotechnical Conference
and 14th ISSMGE Pan-American Conference, Toronto, Canada.
Perla, R., Cheng, T.T., and McClung, D.M., 1980. A two-parameter model of snow avalanche
motion; Journal of Glaciology, v. 26, p. 197-207.
Petley, D., 2012. Remote sensing techniques and landslides. in Landslides: types, mechanisms
and modeling, (ed.) J.J. Clague and D. Stead. Cambridge University Press, Cambridge, p.
159-171.
Peucker, T.K., Fowler, R.J., Little, J.J., and Mark, D.M., 1978. The Triangulated Irregular
Network. Proceedings of the Digital Terrain Models (DTM) Symposium, American
Society of Photogrammetry, p. 516-540.
Pfeiffer, T.J. and Higgins, J.D., 1990. Rockfall Hazard Analysis Using the Colorado Rockfall
Simulation Program; Transportation Research Record, v. 1288, p. 117-126.
Pirulli, M., 2005. Numerical Modelling of Landslide Runout: a Continuum Mechanics
Approach; Ph.D. thesis, Politechnico di Torino, Torino, Italy.
Pirulli, M., Preh, A., Roth, W., Scavia, C., and Poisel, R., 2003. Rock avalanche run out
prediction: combined application of two numerical methods; in Proceedings of the
International Symposium on Rock Mechanics, South African Institute of Mining and
Metallurgy, Johannesburg, p. 903-908.
Pitman, E.B., Nichita, C.C., Patra, A., Bauer, A., Sheridan, M., and Bursik, M., 2003.
Computing granular avalanches and landslides; Physics of Fluids, v. 15, p. 3638-3646.
Poisel, R. and Preh, A., 2008a. Modifications of PFC3D for rock mass fall modeling. in
Continuum and Distinct Element Numerical Modeling in Geo-Engineering - 2008, (ed.) R.
Hart, C. Detournay and P. Cundall, Paper 01-04, Minneapolis, Itasca.
Poisel, R. and Preh, A., 2008b. 3-D landslide runout modelling using the Particle Flow Code
PFC3D; in Proceedings of the 10th International Symposium on Landslides and
Engineered Slopes, ed. Z. Chen, J, Zhang, Z. Li, F. Wu and K. Ho. Taylor and Francis, p.
873-879.
Poisel, R. and Roth, W., 2004. Run out models of rock slope failure; Felsbau, v. 22, p. 46-50.
Porter, M. and Morgenstern, N., 2013. Landslide Risk Evaluation, a Contribution to the
Canadian Technical Guidelines and Best Practices on Landslides; Geological Survey of
Canada, Open File 7312, 21 pages.
Qubec (Province de), 2005. Cartographie des zones exposes aux glissements de terrain dans
les dpts meubles, Guide dutilisation des cartes de zones de contraintes et dapplication
du cadre normative, Saguenay-Lac-Saint-Jean.
Rahman M.G. and Kilgour, D.A., 2000. Gardiner Dam, Three Decades of Performance
Monitoring; in Proceedings of the 2000 Canadian Dam Association Annual Conference,
Regina, SK, p. 212.
RapidEye, 2012. Space Segemtn A constellation of five identical satellites.
http://www.rapideye.com/about/satellites.htm. Accessed 18 September 2012
Reid, M.E. and Brien, D.L., 2006. Assessing massive flank collapse at stratovolcanoe using 3D slope stability analysis; in Landslides from Massive Rock Slope Failure, (ed.) S.G.

56

Evans, G.S. Mugnozza, A.L. Strom, and R.L. Hermanns. Springer, New York, p. 445458.
Revellino, P., Hungr, O., Guadagno, F.M., and Evans, S.G., 2004. Velocity and runout
simulation of destructive debris flows and debris avalanches in pyroclastic deposits,
Campania region, Italy; Environmental Geology, v. 45, p. 295-311.
Rickenmann, D., 1999. Empirical relationships for debris flows; Natural Hazards, v. 19, p. 4777.
Roberson, J.A. and Crowe, C.T. 1993. Engineering Fluid Mechanics, 5th Edition; John Wiley
and Sons, Inc.
Rocscience, 2011. Rocscience Software Products: DIPS V. 5.1, Rocplane V.2, Swedge V. 5,
Slide V.6, Phase2 V.8. Toronto, Ontario, Canada, Rocscience Inc.
[www.rocscience.com].
Sainsbury, B., Pierce, M. E., and Mas Ivars, D., 2008. Analysis of caving behaviour using a
synthetic rock mass-ubiquitous joint rock mass modelling technique. in Proceedings of
First Southern Hemisphere International Rock Mechanics Symposium (SHIRMS 2008),
(ed.) Y. Potvin, J. Carter, A. Dyskin and R. Jeffrey. Australian Centre for Geomechanics,
Nedlands, Australia, p. 243253.
Sarma, S.K., 1973. Stability analysis of embankments and slopes; Geotechnique v. 23, p. 423433.
Sassa, K., 2006. International Society for Rock Mechanics, Commission on Application of
Geophysics to Rock Engineering, suggested method for borehole geophysics in rock
engineering; International Journal of Rock Mechanics and Mining Sciences, v. 43, p.
337-368.
Saunders, W. and Glassey, P. (compilers), 2007. Guidelines for assessing planning policy and
consent requirements for landslide prone land; New Zealand Geological and Nuclear
Sciences Miscellaneous Series 7.
Scheidegger, A.E., 1973. On the prediction of the reach and velocity of catastrophic
landslides; Rock Mechanics, v. 5, p. 231-236.
Schilling, S.P., Griswold, J.P., and Iverson, R.M., 2008. Using LAHARZ to forecast
inundation from lahars, debris flows and rock avalanches: confidence limits on
prediction; in Proceedings of the American Geophysical Union, 2008
Schnabel, P.B., Lysmer, J., and Seed, H.B., 1972. SHAKE: A Continuous Program for
earthquake Response Analysis of Horizontally Layered Sites, University of California,
Berkeley, Earthquake Engineering Research Center, Technical Report UBC/EERC-72/12.
Sitar, N. and MacLaughlin, M. M., 1997. Kinematics and discontinuous deformation analysis
of landslide movement. in 2nd Pan-American Symposium on Landslides, International
Society for Soil Mechanics and Geotechnical Engineering, Rio de Janeiro, Brazil.
Singhroy, V., 2008. Satellite remote sensing applications for landslide detection and
monitoring; in Landslide Disaster Risk Reduction, (ed.) K. Sassa and P. Canuti. Springer,
New York, p. 143-158.
Spencer, E., 1967. A method of analysis of the stability of embankments assuming parallel
inter-slice forces; Geotechnique, v. 17, p. 11-26.
Spratt, D.A. and Lamb, M.A., 2005. Borehole data interpretation and orientation: Turtle
Mountain monitoring project. WP15b Report. University of Calgary, 14 pages.

57

Stead, D., Eberhardt, E., and Coggan, J.S., 2006. Developments in the characterization of
complex rock slope deformation and failure using numerical modelling techniques;
Engineering Geology, v. 83, p. 217-235.
Stead, D. and Coggan, J., 2012. Numerical modeling of rock-slope instability. in Landslides:
Types, Mechanisms and Modeling, (ed.) J.J. Clague and D. Stead, Cambridge University
Press, New York, p. 144-58.
Stevens, W.D., 1998. RocFall: a tool for probabilistic analysis, design of remedial measures
and prediction of rockfalls; M.A.Sc. thesis, University of Toronto, Toronto, Ontario,
Canada.
Takahashi, T. and Yoshida, H., 1979. Study on the deposition of debris flows Part 1:
Deposition due to abrupt change of bed slope. Annals of the Disaster Prevention
Research Institute, Kyoto University, 22B-2. (in Japanese)
Tang, C. and Hudson, J.A., 2010. Rock failure Mechanisms: Illustrated and Explained. CRC
Press, Taylor & Francis Group, New York, 364 pages.
Thompson, S.C., Clague, J.J., and Evans, S.G., 1997. Holocene activity of the Mt. Currie
scarp, Coast Mountains, British Columbia, and implications for its origin; Environmental
and Engineering Geoscience, v. 3, p. 329-348.
VanDine, D., 2011. Professional Practice and Insurance Issues Canadian Technical
Guidelines and Best Practices related to Landslides: a national initiative for loss
reduction. Geological Survey of Canada, Open File 6981, 15 pages.
VanDine, D., 2012. Risk Management Canadian Technical Guidelines and Best Practices
related to Landslides: a national initiative for loss reduction. Geological Survey of
Canada, Open File 6996, 8 pages.
van Westen, C.J., Castellanos, E., and Kuriakose, S.L., 2008. Spatial data for landslide
susceptibility, hazard, and vulnerability assessment: an overview; Engineering Geology,
v. 102, p. 112-131.
Vick, S.G., 2002. Degrees of Belief, Subjective Probability and Engineering Judgment; ASCE
Press, 455 pages.
Wang, C., Tannant, D.D., and Lilly, P.A., 2003. Numerical analysis of the stability of heavily
jointed rock slopes using PFC2D; International Journal of Rock Mechanics and Mining
Sciences, v. 40, p. 415-424.
Watson, A.D., Moore, D.P., and Stewart, T.W., 2004. Temperature influence on rock slope
movements at Checkerboard Creek. in Landslides: Evaluation and Stabilization, (ed.)
W.A. Lacerda, M. Erlich, S.A.B. Fontoura and A.S.F. Sayao. A.A. Balkema, Rotterdam,
p. 12931298.
Wieczorek, G. and Glade T., 2005. Climatic factors influencing triggering of debris flows; in
Debris-flow hazards and related phenomena, (ed.) M. Jakob and O. Hungr.
Praxis/Springer, p. 325-362.
Willenberg, H., 2004. Geologic and kinematic model of a complex landslide in crystalline
rock (Randa, Switzerland). D.Sc. Thesis, Engineering Geology, ETH Zurich,
Switzerland.
Wise, M.P., Moore, G.D., and VanDine, D.F. (editors). 2004. Landslide Risk Case Studies in
Forest Development Planning and Operations; BC Ministry of Forests, Research Branch,
Victoria, BC, Land Management Handbook 56, 119 pages.
Wyllie, D.C. and Mah, C.W., 2004. Rock Slope Engineering: Civil and Mining 4th edition;
Spon Press, 431 pages.
58

Zhang, K., Chen, S.-C., Whitman, D., Shyu, M.-L., Yan, J., and Zhang, C., 2003. A
progressive morphological filter for removing nonground measurements form airborne
LiDAR data; IEEE Transactions on Geoscience and Remote Sensing, v. 40, p. 872-882.

59

You might also like