Você está na página 1de 62

Accepted Manuscript

A critical review of rock slope failure mechanisms: The importance of structural


geology

Doug Stead, Andrea Wolter

PII: S0191-8141(15)00033-4
DOI: 10.1016/j.jsg.2015.02.002
Reference: SG 3184

To appear in: Journal of Structural Geology

Received Date: 31 August 2014


Revised Date: 3 February 2015
Accepted Date: 3 February 2015

Please cite this article as: Stead, D., Wolter, A., A critical review of rock slope failure mechanisms: The
importance of structural geology, Journal of Structural Geology (2015), doi: 10.1016/j.jsg.2015.02.002.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

1 A critical review of rock slope failure mechanisms: The importance


2 of structural geology
3 Doug STEADa and Andrea WOLTERb
a
4 Simon Fraser University, 8888 University Dr., Burnaby, BC, Canada;
5 dstead@sfu.ca
b
6 ETH Zurich, 5 Sonneggstrasse, Zurich, Switzerland; awolter@ethz.ch

PT
7 Abstract

RI
8 Geological structures such as folds, faults, and discontinuities play a critical role in

9 the stability and behaviour of both natural and engineered rock slopes. Although

SC
10 engineering geologists have long recognised the importance of structural geology in

11 slopes, it remains a significant challenge to integrate structural geological mapping

U
12 and theory into all stages of engineering projects. We emphasise the importance of

13
AN
structural geology to slope stability assessments, reviewing how structures control

14 slope failure mechanisms, how engineering geologists measure structures and


M
15 include them in slope stability analyses, and how numerical simulations of slopes

16 incorporate geological structures and processes.


D
TE

17 Keywords rock slopes; stability; failure mechanisms; engineering geology; structural

18 control
EP

19 1. Introduction
C

20 Our understanding of rock slope failure mechanisms has increased considerably


AC

21 during the last decade in response to continued development of urban populations in

22 mountainous areas and to the challenges faced by engineers and geoscientists in

23 the exploitation of large open pit mines. Design in such large-scale rock masses

24 necessitates consideration of structural geology from the micro-scale to the regional

25 tectonic scale. In this paper, we critically review rock slope failure mechanisms with

26 an emphasis on how knowledge of the structural geological environment influences


ACCEPTED MANUSCRIPT

27 all stages in the slope characterisation or design process. We discuss the impacts of

28 broad structural feature types on rock slopes, highlighting important aspects using

29 relevant case studies. We then summarise how structures are characterised in

30 engineering geological projects, and finally review how structural geology is

PT
31 incorporated into numerical modelling of slopes, a useful technique in slope design.

32 Our objective is to demonstrate the critical role of structural features and processes

RI
33 in controlling rock slope stability and failure type, style, and mechanism.

SC
34 2. Effects of structural geology on rock slope stability

35 Structural features, such as folds, faults, and discontinuities, control rock mass

U
36 (i.e., intact rock dissected by discontinuities) behaviour and contribute to either the

37
AN
stabilisation or destabilisation of rock slopes, depending on their orientations and the

38 intensity of associated tectonic damage. Glastonbury and Fell (2000) demonstrated


M
39 how the geometry and composition of a rock slope and its structures determine the
D

40 potential mechanism of a landslide, ranging from translational to complex multi-


TE

41 mechanism failure (Figure 1 and Figure 2). Stead et al. (2006) referred to the

42 importance of structure in determining the complexity of failure mechanisms. In this


EP

43 section, we review the importance of tectonic environment and damage and common

44 brittle and ductile structures related broadly to lithology, with examples drawn from
C

45 the published rock slope literature. Table 1 summarises the structural and lithological
AC

46 features of the cited case studies and additional examples. Specific structural

47 features are discussed in relation to each lithological rock type; however, this does

48 not imply they are exclusive to that lithology.

49 2.1. Tectonic environment and damage


ACCEPTED MANUSCRIPT

50 The tectonic environment and history, or inheritance, of a given slope can

51 determine if and how it fails; in situ stress conditions are important to consider when

52 assessing rock mass behaviour. Hoek et al. (2009) discussed the role of in situ

53 stresses in influencing the design of large open pits. They demonstrated numerically

PT
54 that in situ stresses are typically not significant as compared to uncertainties related

55 to the geological model, strength and deformation properties, and groundwater

RI
56 pressures within the rock masses of a given pit slope. Thus, gravitational stress

SC
57 usually dominates slope stability in open pits. The exception is in areas of high

58 horizontal stress (compressional regime), where stresses can influence open pit

U
59 performance and should be considered. Kinakin and Stead (2005) demonstrated the

60

61
AN
importance of in situ stress conditions in the formation of sackung-type slope failures

(or Deep-Seated Gravitational Slope Deformations (DSGSDs)). Also investigating


M
62 natural slopes, Ambrosi and Crosta (2011) emphasised the interaction between

63 geomorphological processes and in situ stress conditions and slope geometry. They
D

64 demonstrated that these factors influence slope failure type and behaviour. The
TE

65 World Stress Map (Zoback, 1992) is a useful tool in assessing general in situ stress

66 conditions.
EP

67 Past and present seismicity is also important to slope stability. Not only does each

68 earthquake have the capability of triggering slope failures, but the cumulative effect
C

69 of regional seismicity may also damage and weaken slopes. Moore et al. (2011), for
AC

70 example, illustrated the production of slope amplification effects at the 1991 Randa

71 instability in Switzerland due to seismicity. Damjanac et al. (2013) discussed the

72 simulated stability of large open pits in relation to seismic activity and showed that

73 open pits are less susceptible to seismically induced failure than natural slopes.

74 Brian et al. (2014) examined the role of microseismicity in producing fatigue and
ACCEPTED MANUSCRIPT

75 damage in slopes, concluding that microseismic damage is spatially and temporally

76 limited and may be episodic; it typically also only occurs in areas of pre-existing

77 damage (Gischig et al., submitted; Wolter et al., submitted).

78 Numerous constitutive failure criteria have been developed to represent the

PT
79 deformation and strength of rock masses in compression and extension. The most

RI
80 common criterion in rock slope analysis remains the simple, linear Mohr-Coulomb

81 envelope. The Hoek-Brown envelope, originally developed by Hoek and Brown

SC
82 (1980), is also commonly applied to represent the strength of rock masses. Recently,

83 Diederichs (2003) developed a trilinear failure criterion, illustrating the importance of

U
84 brittle fracture through tensile failure and spalling at low confinement. Leith et al.

85 AN
(2014a and b) applied this criterion to two Alpine valleys to investigate the effects of

86 geomorphic and tectonic processes on fracture initiation and propagation, and


M
87 determined stress paths to failure given in situ stress and landscape evolution. Stead

88 and Eberhardt (2013) highlighted the relevance of a trilinear failure envelope to rock
D

89 slope damage.
TE

90 A considerable amount of research in structural geology has been undertaken on


EP

91 damage associated with geological structures (e.g., Shipton and Cowie, 2003; 2008).

92 A rock slope may be subjected to many forms of damage including tectonic damage,
C

93 gravitational displacement-induced damage, geomorphic damage, seismic damage,


AC

94 and damage due to alteration or weathering. Only recently has the importance of

95 rock damage been fully realised in the rock slope stability community; much of this

96 has been due to development of both rock slope mapping technologies and

97 geomechanical modelling incorporating brittle fracture. Brideau et al. (2009) and

98 Brideau (2010) discussed the importance of damage in rock slope instability with

99 reference to several major landslides including the 1965 Hope, 2002 McAuley Creek,
ACCEPTED MANUSCRIPT

100 and 2007 Chehalis Lake landslides in British Columbia, and the South Peak, Turtle

101 Mountain instability in Alberta, Canada. Pedrazzini et al. (2011) demonstrated the

102 varied rock mass damage associated with folding at Turtle Mountain and its effect on

103 instability. More recently, Agliardi et al. (2013) and Stead and Eberhardt (2013)

PT
104 provided more in-depth discussions on the importance of rock mass damage to

105 slope stability. We emphasise the importance of tectonic damage on rock masses.

RI
106 This damage reduces the strength of rock masses and also provides the kinematics

SC
107 required for slope displacements; we suggest that tectonically induced damage has

108 a fundamental influence on subsequent gravitational damage and is interlinked with

U
109 other forms of damage.

110 AN
2.2. Importance of structures in sedimentary rock slopes

111 Sedimentary rock masses typically behave anisotropically due to bedding planes,
M
112 which act as planes of weakness along which slip may occur. Closely spaced
D

113 bedding planes dipping into the slope may lead to toppling slope failures. Such
TE

114 failures are described in surface coal mines, where repeated strata of arenaceous

115 and argillaceous rock masses dip into the slope (Benko, 1997). Stability of these
EP

116 slopes is controlled by the relative spacing of bedding and cross joints, the presence

117 of lateral release joints, and the interbed shear strength. The influence of tectonic
C

118 activity may lead to a reduction in the interbed shear strength with an important
AC

119 influence on both sliding and toppling movements.

120 Originally horizontal beds are frequently uplifted, tilted, and folded. The folding of

121 sedimentary rock sequences has numerous and varied effects on both natural and

122 open pit slope stability (Figure 3). The orientation of the topographic surface or

123 excavation face with respect to bedding is of major significance. Knowledge of the
ACCEPTED MANUSCRIPT

124 fold type and orientation, and its structural complexity, are important in realistic rock

125 slope modelling. We suggest that the features of a conventional fold classification,

126 such as Fleuty (1964) or Ramsay (1967), should be considered with respect to how

127 they influence rock slope stability (Figure 4). Horizontal folds may be associated with

PT
128 simple translational failure (sliding of rock blocks in the true dip direction) on one limb

129 and toppling mechanisms on the other fold limb, depending on their orientation with

RI
130 respect to a given slope face (Figure 4i-iii). Steeply plunging folds (Figure 4iv-vii)

SC
131 may result in complex translational and rotational slope displacements in both planar

132 and toppling modes (Brideau, 2005; Yan, 2008). The bluntness and tightness of a

U
133 fold will not only influence the kinematics of slope failure but may also be indicative

134

135
AN
of the degree of tectonic damage (reduced rock mass quality) sustained by the rock

mass in the form of both fracturing and interbed slip (Figure 5). Open folds may show
M
136 the typical transverse, longitudinal, and conjugate discontinuities associated with

137 folding, whereas tight and isoclinal folds may develop foliation and cleavage, a sign
D

138 of increased damage. Whether the folding is of a parallel or concentric nature or


TE

139 similar/shear type will also influence kinematics and potential failure block geometry.

140 It is necessary to consider the number of deformation episodes and their


EP

141 combined effect on rock mass kinematics and strength. Folded rock units with

142 plunging hinge lines, for example, have likely sustained more deformation events
C

143 than units with horizontal hinges. Folding in two directions and the presence of
AC

144 interference folds has been shown to influence the complexity of slope deformations

145 within landslides and, in combination with natural release surfaces, may explain

146 commonly observed multiple block-type movement. Figure 6, for example, shows the

147 sliding scar morphology of the 1963 Vajont Slide in Italy. This surface is affected not

148 only by two generations of regional folds, producing complex, metre- to decametre-
ACCEPTED MANUSCRIPT

149 scale dome-and-basin to crescent-and-mushroom 1-2 (Ramsay, 1967) or K

150 (Thiessen and Means, 1980) type interference patterns (Figure 6b; Massironi et al.,

151 2013), but also by concretions on the centimetre scale. Folding plays a vital role in

152 providing what are referred to as geologic release surfaces for rock slope failures on

PT
153 the sides (lateral release), at the back (rear release), and on the base (basal plane)

154 of failure blocks. Discontinuity sets associated with folding (i.e. transverse,

RI
155 longitudinal and conjugate) may play a major role in releasing the blocks that slide

SC
156 along folded bedding surfaces – such release surfaces provide the necessary

157 kinematic freedom for slope failure. The association of kinematic rock slope failure

U
158 modes with anticlinal folding has been discussed by Badger (2002), who described

159

160
AN
the relationship between failure (planar/toppling) and the relative location of the rock

slope and discontinuity sets commonly associated with folded, bedded units. Figure
M
161 7 illustrates how three possible configurations of the folded bedding (S0) or primary

162 discontinuity set (S1) and secondary discontinuity set (S2) can produce different
D

163 sliding failures. Another example of structural analysis associated with natural slope
TE

164 failures is that conducted by Humair et al. (2013) on the 1903 Frank Slide area in

165 Alberta, Canada (Figure 8); they demonstrated the importance of the Turtle Mountain
EP

166 Anticline orientation, dimensions, and associated fractures to slope stability. Relative

167 slope and fold orientations, as well as damage intensity, affect slope stability and the
C

168 range in complexity of the failure kinematics.


AC

169 Fault planes often form sliding surfaces or release surfaces and can be associated

170 with the steepening of bedding (drag folding) that may induce slope instability. For

171 example, Figure 9a shows a failure in a coal mine, where a fault acted as a rear

172 release plane for a translational failure that occurred along drag-folded strata

173 adjacent to the fault. Figure 9b shows the fault exposed within the mine and the drag
ACCEPTED MANUSCRIPT

174 folding. As discussed previously, damage zones associated with faults may similarly

175 result in a reduction in rock mass quality influencing slope stability. Limited work in

176 rock slope engineering has been undertaken to date considering fault contact

177 relationships and their influence on rock slope instability; this includes not only

PT
178 damage but also fault linkage, drag mechanisms, fault seals, shale smear, the nature

179 of fault gouge, cataclasis, and diagenetic effects.

RI
180 Roughness exerts a primary control on rock slope failures and exists at all scales

SC
181 from primary and secondary small-scale asperities to large-scale undulations. Focus

182 in rock engineering has been on the use of roughness measures such as the Joint

U
183 Roughness Coefficient, JRC (Barton 1976, 2011; Barton and Choubey 1977). These

184 AN
roughness measures have been linked to empirical approaches for assigning shear

185 strength to joints. The JRC, for example, is used in the following equation:
M
tan log
D

186 where τ is shear strength along the discontinuity plane, σn is the normal stress on the
TE

187 plane, φr is the residual friction angle, and JCS is the Joint Compressive Strength.
EP

188 The importance of sedimentary and tectonic structures on discontinuity surfaces has

189 in the authors’ opinions received insufficient attention. Tectonic effects on


C

190 roughness, such as slickensides and polishing, have been recognised as indicators
AC

191 of residual strength and the need for testing the shear strength in the appropriate

192 direction with respect to surface structures noted. Cruden and Krahn (1978)

193 described the presence of slickensides on folded bedding surfaces, as well as minor

194 thrusts, which controlled the Frank Slide.


ACCEPTED MANUSCRIPT

195 We suggest that sedimentary structures, such as sole marks, flute casts,

196 concretions, and ripple marks, also may have a significant influence on local or

197 micro- to meso-scale roughness, and hence affect global rock slope behaviour. For

198 example, concretions or ripple marks may increase bedding plane frictional

PT
199 resistance and in some cases even induce dilation and the need for shearing

200 through asperities during failure. Such features may act as rock bridges along

RI
201 potential sliding surfaces. Rock bridges are zones of intact rock classified as either i)

SC
202 in-plane rock bridges on a discontinuity surface, or ii) out-of-plane rock bridges

203 separating individual discontinuities (Tuckey, 2012).

U
204 2.3. Importance of structures in igneous rock slopes

205
AN
Rock slope failures in igneous rock types are highly varied, depending on the

206 topography, in situ stress conditions, and degree of weathering or alteration. Given
M
207 the high strength of granites and similar igneous rock masses, it is not surprising that
D

208 it is invariably the structures within the igneous rock that form the weak link and
TE

209 control slope instability. One example of structural controls in igneous rock types is

210 the effect of sheet, or exfoliation, joints. Hencher and Richards (1982), Hencher et al.
EP

211 (2011), and Ziegler et al. (2013, 2014) discussed the importance of sheet joints,

212 proto-joints, persistence, rock bridges and fractography in engineering geology,


C

213 focussing on joint initiation and propagation in slopes, as well as their effects on
AC

214 slope stability and determining in situ stresses from fracture patterns. As sheet joints

215 generally parallel the topography, instability may involve undercutting by excavation,

216 failure by linkage with other structures and, in the case of high-stress environments,

217 the fracturing and development of breakout-type failures.


ACCEPTED MANUSCRIPT

218 When considering structurally controlled stability of igneous rock slopes the

219 importance of discontinuity persistence in addition to orientation and intensity

220 (number of discontinuities along a line, in an area, or in a volume) is emphasised.

221 Non-persistent structures must be taken into account as the strength of rock bridges

PT
222 between discontinuities is high and therefore even small percentages of rock bridges

223 will increase the overall strength of the slope resulting in a higher factor of safety.

RI
224 Consideration should be given to step-path-type failures involving blocks moving on

SC
225 multiple fractures with possible intact brittle rock fracture. Characterisation of fracture

226 networks in igneous rocks for use in subsequent numerical models (see below)

U
227 requires considerable care and often the combined use of field mapping, remote

228 sensing and borehole methods.


AN
229 Fault zones within igneous rock masses act as weak zones within otherwise high
M
230 quality rock, and are thus significant in controlling failure location and behaviour.

231 They also influence groundwater flow, which is an important factor in the alteration of
D

232 igneous rock masses as well as in pore pressure conditions in slopes. At Diavik
TE

233 Diamond Mines in Northwest Territories, Canada, for example, a major fault zone

234 dissects two kimberlite pipes and directs groundwater flow into the existing open pit
EP

235 (Tuckey, 2012; Vivas Becerra, 2014). Considerable work has been undertaken

236 recognising the influence of structure on hydrogeology and pore water pressure at
C

237 Diavik, as described by Milmo et al. (2014).


AC

238 Although usually not a dominant factor in slope stability of igneous rock masses,

239 micro-scale features should also be considered. The different strengths of various

240 minerals included in a particular rock mass, as well as grain and lattice distortions

241 and deformations, affect rock mass weathering and strength. Stress-induced fracture

242 propagation and coalescence in high slopes may ultimately form persistent
ACCEPTED MANUSCRIPT

243 structures along which sliding can occur. Considerable research is ongoing,

244 examining the influence of microstructure heterogeneity and micro-cracking on rock

245 strength (Hamdi et al., 2013; Nicksiar and Martin, 2014).

246 2.4. Importance of structures in metamorphic rock slopes

PT
247 Numerous large, catastrophic rock slope failures have occurred in metamorphic

RI
248 lithologies. Foliations, cleavage, and other deformational features strongly influence

249 stability; the metamorphic facies level of a given rock mass can also have a

SC
250 significant influence on failure by altering rock mass strength. Catastrophic failures

251 have occurred in both high-grade metamorphic rocks such as gneisses and in highly

U
252 foliated slates, phyllites, and schist. Dominant foliations can be parallel, oblique, or

253
AN
opposite to the sliding direction, with mechanisms consequently involving planar

254 sliding, oblique sliding/rotation, or toppling. As discussed previously, it is important to


M
255 consider different structural deformation phases and their influence on foliations.
D

256 Complex folding of foliations in the dip and strike directions of a slope adds
TE

257 complexity to failure kinematics and observed displacements.

258 An example of a major landslide in metamorphic gneiss terrain is the 1991 Randa
EP

259 rockslide sequence in Switzerland (Figure 10) (Eberhardt et al., 2004; Willenberg et

260 al., 2008; Gischig et al., 2009; Löw et al., 2012). This rockslide has been the subject
C

261 of extensive research over the last decade and is an excellent example of the
AC

262 influence of geological structure on rock slope failure. Geological structures form the

263 failure surface and lateral release surfaces (Table 1).

264 Another example is the Åknes rock slope in Norway (Table 1) (Ganerød et al.,

265 2008, Grøneng et al., 2011; Blikra, 2012). This active slope instability is located in a

266 highly deformed metamorphic region comprising foliated schistose rock types on the
ACCEPTED MANUSCRIPT

267 western slope of the Storfjorden fjord and presents a major risk of landslide-induced

268 tsunami. The influence of the metamorphic structures in controlling the active

269 movement of the rock slope has been shown through geophysical and borehole

270 methods, geotechnical instrumentation and numerical modelling. The basal surface

PT
271 of the landslide is stepped, following discontinuities separated by intact rock bridges.

272 An interesting aspect of this slide is the multiple blocks involved in the movement

RI
273 with different rates and movement directions; these may be related to three-

SC
274 dimensional variations in the geological structure along the basal sliding surface.

275 Such movement of blocks is not uncommon and shows the importance of

U
276 understanding three-dimensional geological structure.

277 AN
The Downie Slide in British Columbia, Canada, is the largest prehistoric landslide

278 in North America with an estimated volume of 1.5 billion m3. The toe of this landslide
M
279 was submerged by a reservoir impounded by the Revelstoke Dam, and hence this

280 landslide has been extensively investigated. A major slope dewatering scheme was
D

281 developed to ensure stability; an excellent discussion of the characterisation and


TE

282 modelling of this landslide is presented in Kalenchuk (2010). Like Åknes, this slide

283 comprises multiple blocks and is located in metamorphic mica schist, gneiss, and
EP

284 quartzite.
C

285 Examples of toppling-induced landslides in foliated metamorphic rock types are


AC

286 provided in Newman (2013) and Clayton (2014). An excellent discussion on toppling

287 instability is presented in Goodman (2013). When considering the influence of

288 metamorphic structures on hazards associated with rock slopes, it is important to

289 consider both primary and secondary slope failure mechanisms. A useful example is

290 presented by Brideau et al. (2006) at the East Gate slope in Glacier National Park,

291 British Columbia, Canada. The upper unstable rock slope is situated within phyllites
ACCEPTED MANUSCRIPT

292 and involved pseudo-circular rock mass failure influenced by foliations including

293 crenulation cleavage. Although this failure was at considerable distance from a major

294 highway (Trans-Canada Highway) at the foot of the mountain slope, it led to the

295 gradual accumulation of friable phyllitic failure debris. Over time, this debris has

PT
296 remobilised as debris flows that have travelled downslope and caused significant

297 blockages of the Trans-Canada Highway.

RI
298 Deep-seated gravitational displacements are an important variant of deforming

SC
299 rock slopes, which occur in multiple rock types and are particularly common in post-

300 glacial metamorphic mountain terrains. DSGSD slopes have characteristic

U
301 topographic features such as antiscarps, half-grabens, double-crested ridges,

302 AN
trenches, and slope toe bulging, providing evidence of previous or active movement.

303 Although beyond the scope of the current paper, it is important to emphasise that
M
304 DSGSDs are often located in foliated rock slopes; mechanisms of displacement may

305 be highly variable, including translational sliding, circular and toppling failures. The
D

306 La Clapière landslide in France, for example, is part of a larger DSGSD complex
TE

307 (Figure 11) and has shown varied behaviour over the past century (Vengeon et al.,

308 1999; Cappa et al., 2004; Lebourg et al., 2010).


EP

309 3. Engineering geology methods used to characterise structures


C

310 Various techniques have been applied to rock slopes to characterise and analyse
AC

311 them, ranging from field surveys to remote sensing. Traditional rock slope

312 investigations often commence with field observations of the rock mass, including

313 intact rock descriptions, and discontinuity measurements and characterisation.

314 Lithology, degree of weathering, and intact rock strength are determined through

315 observation and in situ or laboratory tests. Stress conditions may be measured on
ACCEPTED MANUSCRIPT

316 site using overcoring, flat jack, or hydraulic fracturing tests to provide direction and

317 magnitude of stress, or using earthquake focal mechanism analysis, breakout and

318 failure plane orientation observations, or stress relief measurements to determine the

319 direction of stress. Measurement of in situ stress conditions in slopes is nonetheless

PT
320 rare and remains a challenge, as a wide variation in results are common.

RI
321 Discontinuity surveys using line surveys or circular/rectangular windows allow the

322 definition of discontinuity sets; the measurement and assessment of discontinuity

SC
323 characteristics, such as persistence, intensity, roughness, aperture, infill and

324 seepage; and the delineation of structural domains. Recently, Hencher (2013) stated

U
325 that, although the method of discontinuity surveying is simple and straightforward,

326 AN
the definition of discontinuities is not. He suggested a revised classification scheme

327 of discontinuities that includes an estimation of their tensile strength. This indicates
M
328 explicitly that not all discontinuities have the same effect on rock mass strength and

329 slope stability.


D
TE

330 By considering specific slope (or design sector) orientations, the potential stability

331 of a rock slope may be investigated using stereographic-based kinematic analyses.


EP

332 These methods have changed little since their introduction, and are summarised in a

333 large number of publications including Richards et al. (1978), Priest (1980), Hoek
C

334 and Bray (1981), and Wyllie and Mah (2004). Computer codes such as DIPS
AC

335 (Rocscience, 2014) are now routinely used in kinematic analysis of rock slopes to

336 assess potential simple planar, wedge and toppling (flexural and direct) instability

337 modes (Figure 12). Limitations of this method can include insufficient consideration

338 of discontinuity persistence, spacing, and other factors controlling discontinuity shear

339 strength. Over-reliance on contoured discontinuity orientation data alone can

340 produce either conservative or, more importantly, unsafe stability assessments. A
ACCEPTED MANUSCRIPT

341 recommended procedure is to consider first the influence of major discrete

342 structures, such as faults/shear zones and persistent bedding planes, on slope

343 stability and then incorporate instability mechanisms related to joint sets. Ground-

344 truthing of stereographic kinematic analyses through field investigations is strongly

PT
345 suggested. Most kinematic analyses are restricted to simple planar, toppling or

346 tetrahedral (+/- tension crack) geometries. Often potential failures will have more

RI
347 complex geometries involving multiple basal, rear and lateral release surfaces. As an

SC
348 example, the truncation of a tetrahedral wedge by basal surfaces such as bedding

349 planes often results in a pentahedral or hexahedral wedge geometry. Kinematic

U
350 analysis of rock slopes, with the exception of direct toppling, is also generally limited

351

352
AN
to consideration of translational failure mechanisms with limited consideration of

rotational moments or intact rock fracture (Hungr and Amann, 2011). The most
M
353 recent versions of rock slope wedge analysis programs (e.g., Swedge (Rocscience,

354 2014)) allow for limit equilibrium analysis of pentahedral wedge geometries. Figure
D

355 13 indicates two possible geometries, each with basal and lateral release planes and
TE

356 Figure 13b with a rear tension crack. The use of kinematic analysis methods in slope

357 stability assessments are suggested to be in general more suited to surface mine
EP

358 bench scale or highway road cuts than to large-scale landslides or open pit mine

359 slopes unless high-persistence, discrete structures control instability.


C

360 A rapidly developing number of remote sensing techniques, including laser


AC

361 scanning, digital photogrammetry, InSAR and thermal imaging are allowing, in

362 combination with conventional field mapping, improved characterisation of rock

363 slopes (Jaboyedoff et al. (2007), Lato et al. (2009), Sturzenegger and Stead (2009),

364 and Petley (2012)). Using such ground-based, airborne and satellite-based methods

365 provides the ability to map previously inaccessible rock slope areas, often with sub-
ACCEPTED MANUSCRIPT

366 centimetre accuracy. A major recent development, increasingly used in engineering

367 geological mapping, is the use of Unmanned Airborne Vehicle (UAV) platforms.

368 Remote discontinuity surveys allow the measurement of discontinuity orientation,

369 persistence, spacing, and even discontinuity roughness (see, for example Haneberg,

PT
370 2007; Tatone and Grasselli, 2010); ductile structural features such as folds may also

371 be characterised. As demonstrated by Wolter et al. (2014), fold hinge orientations

RI
372 and interference patterns may be quantitatively assessed using remote sensing, with

SC
373 implications for failure initiation and propagation. If possible, field validation of remote

374 observations is advised.

U
375 Finally, subsurface studies, where conducted, contribute greatly to the improved

376 AN
spatial characterisation of rock slope masses, and may include both borehole

377 logging and geophysical surveys. For example, geophysical borehole logging, 1D
M
378 borehole radar reflection surveys, 3D surface georadar, and seismic refraction

379 surveys, as well as inclinometer and televiewer data were all used in the
D

380 characterisation of subsurface structures at the Randa slope (Löw et al., 2012).
TE

381 Clayton (2014) showed the use of borehole Time-Domain Reflectometry (TDR) in

382 characterising the location of landslide failure surfaces.


EP

383 One of the most extensively and intensively monitored sites is the Åknes slope
C

384 (Grøneng et al., 2011; Blikra, 2012). Here, borehole data sets, such as core logs,
AC

385 measurements of water pressure and flow and penetration velocity, resistivity

386 profiles, and P-wave velocity profiles, have allowed the determination of hydrologic

387 conditions and rock mass quality, and the identification of large fractures and sliding

388 surfaces underground. These data complement surface observations and

389 measurements of structures such as discontinuities. Unfortunately, for many natural


ACCEPTED MANUSCRIPT

390 rock slope instabilities in remote mountainous areas such data are rare and often

391 prohibitively expensive to obtain.

392 4. Structural geology and the numerical modelling of rock slopes

PT
393 4.1. Limit equilibrium analysis

394 The simplest and most common form of slope analysis involves force and moment

RI
395 equilibrium and uses iterative methods of slices to calculate the factor of safety (FS =

SC
396 resisting forces (or moments) / driving forces (or moments)). The most accurate

397 methods determine the factor of safety for both moment and force equilibrium and

U
398 use different assumptions for the interslice forces (Bromhead, 1992; Duncan and

399

400
AN
Wright, 2005). The methods incorporate a searching algorithm that, in its simplest

form, determines the critical circular slip surface for a slope (lowest factor of safety),
M
401 which is in reality only valid for extremely weak rock masses or soil-like materials. To

402 allow analyses of more complex slope failure mechanisms, composite or block
D

403 searching algorithms have been developed that search for critical, non-circular block
TE

404 geometries. Facilities exist to import groundwater conditions varying from simple

405 water tables to coupled limit equilibrium-groundwater numerical models, thereby


EP

406 providing the ability to consider transient slope instability. Seismic disturbance may

407 be included using a pseudo-static approach and a wide range of support may be
C

408 applied in the software. In practice the influence of geological structure often
AC

409 determines the location of the critical sliding surface which may be:

410 • Along major discrete structures, such as bedding planes (usually adversely

411 dipping or folded), faults, and shear zones. These may lead to single-plane

412 or multi-plane (e.g. combining bedding and faults) failures.


ACCEPTED MANUSCRIPT

413 • Through the rock mass and controlled by anisotropy within the rock mass

414 due to discontinuity sets (bedding, jointing, cleavage, foliation).

415 • A combination of failure along discrete structures and failure through intact

416 rock.

PT
417 Recent developments have seen an increasing use of anisotropic functions in limit

RI
418 equilibrium analyses to allow for directional weakness created by pre-existing

419 geological structure (Rocscience, 2014). Modifications have also allowed the

SC
420 inclusion of rock bridges by the use of apparent cohesion and friction values to

421 represent percentage of intact rock that must be sheared through for failure to occur

U
422 (Jennings, 1970; Baczynski, 2000). Analyses have in the past been predominantly

423 AN
deterministic, providing one factor of safety for the slope, and have often been

424 accompanied by sensitivity analyses to determine the influence of geological


M
425 structure, slope geometry, mechanical properties and groundwater. Both two- and

426 three-dimensional limit equilibrium methods exist; however three-dimensional


D

427 methods are mostly applied in the analysis of simple tetrahedral wedge geometries.
TE

428 Elegant methods of columns have been developed to allow three-dimensional limit

429 equilibrium analyses of more complex failure surfaces, but have seen less frequent
EP

430 use in the analysis of landslides. Probabilistic analyses have become increasingly

431 common, with the need to consider the risk and consequence of slope instability and
C

432 both model and parameter uncertainty. Although limit equilibrium methods provide
AC

433 useful preliminary analysis and may be adequate for simple slope failure

434 mechanisms, they consider only forces or moments and not slope displacements,

435 and as such are unable to fully capture complex slope failure mechanisms.

436 4.2. Continuum analysis of rock slope failure


ACCEPTED MANUSCRIPT

437 Numerical modelling of rock slopes is often undertaken assuming a continuum

438 approach, where the rock mass is treated as an equivalent medium. The rock mass

439 is usually meshed using elements of triangular or quadrilateral shape in 2D. A

440 constitutive criterion allows simulation of varied stress-strain behaviour from simple

PT
441 elastic to elasto-plastic or time-dependent creep. The strength of the modelled rock

442 mass comprises components of intact rock and discontinuity strength in

RI
443 compression, shear and tension. The strength is hence inherently scale-dependent

SC
444 and it is here that the influence of structure is paramount. A Geological Strength

445 Index (GSI) ranging from 0 to 100 (Figure 14) was introduced by Hoek et al. (1995)

U
446 and modified by Cai et al. (2004) and Hoek et al. (2013), and is currently used as the

447

448
AN
main method for deriving equivalent media properties of a rock mass suitable for

incorporation into 2D and 3D continuum geomechanical models. Discontinuity


M
449 spacing (block size) and surface condition are considered to derive the GSI value

450 and, indirectly, the rock mass strength; both of these input criteria are controlled by
D

451 the structural geology of an area. A modification in the GSI approach was proposed
TE

452 for tectonically disturbed flysch rock masses (Marinos and Hoek, 2000) (Figure 15).

453 It is essential always to consider a slope rock mass using a structural geology
EP

454 domain approach recognizing characteristic structures within an area; such a domain

455 approach is often extended to consider hydrogeological and geotechnical domains,


C

456 both again often a function of the tectonic history and structural geology. In open pit
AC

457 mining geomechanics, the approach adopted is to consider the structural geological

458 model of the slope, the hydrogeological model, and the rock mass model. These

459 models are combined in a geotechnical model of the slope or open pit (Read and

460 Stacey, 2009). An adequate structural model of the pit slope is hence an essential

461 prerequisite for optimal slope design.


ACCEPTED MANUSCRIPT

462 The most common continuum numerical modelling approaches for rock slope

463 analysis are the finite element and the finite difference methods. The most popular

464 codes for each approach are, respectively, Phase2 and RS3 (Rocscience, 2014),

465 and FLAC2D and FLAC3D (Itasca, 2014). Both finite element and finite difference

PT
466 methods have the ability to incorporate discrete structures such as faults but these

467 methods are not optimal in the analysis of jointed media. Recent developments in

RI
468 the finite element code Phase2 have allowed incorporation of jointed media (e.g.

SC
469 bedding, discrete fracture networks and Voronoi tessellation) and the modelling of

470 structurally controlled rock slope failures, such as toppling, planar and biplanar

U
471 modes. Figure 16 shows examples of the use of 2D and 3D continuum codes in the

472

473
AN
modelling of landslides and rock slopes. Where discrete major structures control rock

slope instability, the use of the GSI in deriving the rock mass strength is not
M
474 recommended as the shear strength of the structures, not the rock mass, will control

475 the kinematics of slope failure. In such slopes, discontinuum methods are used to
D

476 simulate slope deformation.


TE

477 4.3. Discontinuum analyses

478 To accurately represent jointed media and large displacements of rock slopes,
EP

479 the use of discontinuum simulations is most appropriate. A range of discontinuum


C

480 numerical modelling methods exist and are usually grouped under “Discrete Element
AC

481 Methods”. The most commonly used in rock engineering are the Distinct Element

482 and Discontinuous Deformation Analysis methods. Stead and Coggan (2012)

483 provide a more detailed description of these methods; here, we will focus on the

484 Distinct Element Method, which is the most commonly used for modelling rock

485 slopes. A rock slope is discretized into joint-bounded blocks; the joints are assigned

486 a shear strength constitutive criterion and shear and normal stiffnesses (controlling
ACCEPTED MANUSCRIPT

487 displacement along joints). The intact rock blocks can be rigid or deformable. If the

488 blocks are treated as deformable they can be assigned a wide range of constitutive

489 behaviour including elastic, elasto-plastic, strain-softening, and time-dependent

490 options. Modelling of rock slopes using Distinct Element methods can thus simulate

PT
491 movement along major structures, such as faults, as well as the translation and

492 rotation of joint-bounded blocks, in 2D and 3D. Although sophisticated software

RI
493 exists, considerable challenges remain in simulating the influence of geological

SC
494 structures on rock slope deformation, including:

495 • The persistence of structures, such as joints, both along strike and dip is

U
496 usually difficult to quantify in the field. Most models routinely assume

497 AN
persistent, through-going structures – an over-simplification in practice.

498 • Where non-persistent structures exist, the engineer must consider rock
M
499 bridges.

500 • Rock masses are typically heterogeneous. Properties such as strength and
D

501 discontinuity intensity vary temporally and spatially. Shear zones, damage
TE

502 zones, and folds should be noted and treated as separate domains in

503 analysis.
EP

504 • The level of detail to include is challenging to determine. Although it is

505 impossible to model every discontinuity, engineering judgement must be used


C

506 to determine which structures are important for the specific problem at hand.
AC

507 Scale effects must be considered.

508 Four levels of discontinuum modelling of rock slopes, from the simplest to the

509 most complex, can be recognised in the literature:


ACCEPTED MANUSCRIPT

510 i. Consideration of the dominant discrete failure surfaces only. Figure 17a

511 shows an example application of the 2D UDEC code (Itasca, 2014) in

512 modelling the Vajont landslide. The discrete surface in this analysis follows

513 folded bedding with a low-angle seat and high-angle rear forming a chair-

PT
514 shaped failure surface. Such an analysis allows identification of the

515 interaction mechanisms between the upper driving block and the lower

RI
516 passive block, with the formation of a transitional Prandtl prism or yield

SC
517 zone. Figure 17b shows a similar approach in the 3D Distinct Element

518 code 3DEC (Itasca, 2014), applied to the Beauregard landslide in Italy

U
519 (Kalenchuk, 2010). This model involved consideration of the primary basal

520

521
AN
failure surface, and the lateral and rear release surfaces. It is important to

constrain the failure surface based on observed structures; this was


M
522 possible at Beauregard using available borehole and surface mapping

523 data as well as interpretation of best-fit failure geometry. A comparable 3D


D

524 model of the Mitchell Creek landslide in British Columbia has recently
TE

525 been constructed by Clayton (2014) utilizing detailed borehole

526 investigations and engineering geological / geomorphological surface


EP

527 mapping. It is essential in any 3D geomechanical modelling that engineers

528 work closely with structural geologists in order to optimise the


C

529 interpretation of failure surface geometry and the influence of tectonics on


AC

530 rock mass strength and failure mechanism.

531 ii. Incorporation of the full geological structure – that is, inclusion of both

532 major structures (faults and folds) and discontinuity sets (bedding and joint

533 sets). Joint sets within Distinct Element models are simulated by defining

534 their dip (and dip direction if in 3D), spacing, and strength properties (e.g.
ACCEPTED MANUSCRIPT

535 cohesion, friction angle, dilation, tensile strength). It is also necessary in

536 bedded materials to define an offset between orthogonal joints. The

537 overriding assumption in conventional models is that discontinuities are

538 persistent throughout the model (Figure 16a shows the different

PT
539 assumptions between continuous and discontinuous joints). In order to

540 address variability, it is possible to define a standard deviation for input

RI
541 values such as dip, dip direction, joint stiffness, cohesion, friction angle,

SC
542 and dilation angle. Complex, continuously folded bedding surfaces can be

543 simulated through the use of arc commands as seen in the simulation of

U
544 the Frank Slide in Figure 17c (Benko and Stead, 1998); this figure shows

545

546
AN
the incorporation of two joint sets, discrete structures, and a major fold. An

analysis assuming continuous through-going structures is usually a very


M
547 conservative assumption, as discontinuities are often separated by rock

548 bridges.
D

549 iii. Inclusion of rock bridges along discontinuity surfaces. Rock bridges are
TE

550 incorporated into models by using continuous joints but specifying a

551 persistence factor between 0 and 1, which assigns percentages of rock


EP

552 bridge strength to the overall strength of the discontinuity (for example,

553 Franz (2009); Brideau et al. (2012)). Variants of this approach are used in
C

554 many model types whereby the strength of the discontinuity is adjusted to
AC

555 include a component of strength, simulating non-persistence or rock

556 bridges.

557 iv. Explicit incorporation of rock bridges using a Discrete Fracture Network

558 (DFN). DFNs are increasingly being used in all fields of rock engineering,

559 including underground and surface rock mechanics, petroleum


ACCEPTED MANUSCRIPT

560 geomechanics, enhanced geothermal systems, and groundwater flow in

561 fractured rock. DFN generation can be undertaken using numerous

562 commercial and public domain codes such as Fracman (Golder

563 Associates, 2012), Move4D (Midland Valley, 2014), and FracSim3D (Xu

PT
564 and Dowd, 2010). The principal inputs required from structural mapping for

565 DFN generation are statistical data on joint trace length, joint intensity,

RI
566 joint orientation, and joint termination. Discrete fracture networks are then

SC
567 imported into geomechanical models to allow explicit representation of

568 non-persistent discontinuities and rock bridges. Once again it is extremely

U
569 important to consider the structural geology and tectonic history of an area

570

571
AN
to ensure realistic geomechanical models. It is possible to incorporate

major observed discontinuities into a model and combine these with a


M
572 statistically generated DFN; ground-truthing DFN models is important

573 wherever feasible and care must always be taken to ensure that
D

574 engineering geological judgement is used prior to the simulation of Distinct


TE

575 Element models incorporating DFNs. We emphasise that further research

576 is required in rock engineering in collaboration with structural geologists to


EP

577 characterise discontinuity persistence, rock bridges and termination

578 modes. Characterising the percentage of rock bridges is an important


C

579 challenge of major importance in future rock engineering design. Although


AC

580 this subject has been extensively researched, insufficient attention has

581 been given to the structural geological aspects. Tuckey (2012) and Tuckey

582 et al. (2013) described recent attempts in characterising persistence and

583 rock bridges in large open pit rock slopes. Dowd et al. (2007) and

584 Alghalandis et al. (2014) provided important contributions to understanding


ACCEPTED MANUSCRIPT

585 the fundamental properties of DFNs, including the connectivity of fracture

586 networks. Work such as the Leeds Rock Fracture Experiment on fracture

587 connectivity has important relevance to rock slopes and landslides.

588 Hencher (2013) and Hencher and Richards (2014) described key research

PT
589 in the constraint of fracture geometry in rock masses. Their work

590 emphasised the need to consider the fundamentals of joint genesis in rock

RI
591 engineering investigations with associated implications for geomechanical

SC
592 modelling.

593 4.4. Modelling rock damage and brittle fracture

U
594 Recent developments in geomechanical codes have provided an impetus for

595
AN
consideration of the roles of brittle fracture and damage in large rock slopes.

596 Continuum and discontinuum codes are able to simulate damage in different ways.
M
597 In continuum codes, damage within a rock slope may be characterised implicitly by
D

598 considering the change in number of yielded elements compared to the total number
TE

599 of elements. In discontinuum codes, brittle fracture and damage may be simulated

600 explicitly using a variety of approaches (Table 2). Methods include hybrid Finite-
EP

601 Discrete Element codes (FDEM), Distinct Element codes (DEM) incorporating a

602 Voronoi or Trigon mesh, Particle Flow Codes (PFC), and lattice spring codes. In
C

603 FDEM approaches, the rock slope is discretized into finite elements and, by
AC

604 incorporating a fracture mechanics criterion, the mesh is able to simulate

605 fragmentation from an intact continuum to a discontinuum comprising rock

606 fragments, enabling the full failure process to be captured. The simulation

607 incorporates continuous remeshing of fragments as they are formed by new

608 fracturing. It is also possible to incorporate non-persistent joints into the rock slope

609 either as step-path features or DFNs. Figure 18a shows an example of an open pit
ACCEPTED MANUSCRIPT

610 slope where fracturing of rock bridges within a DFN are induced due to stresses

611 related to underground mining (Vyazmensky et al., 2010).

612 Distinct Element codes may be used to model brittle fracture in both 2D and 3D

613 using four methods – Voronoi, Trigon, PFC, and lattice spring methods. In the first

PT
614 method, intact rock properties are given to the sides of the polygonal Voronoi mesh;

RI
615 cracks develop when the induced stresses cause failure of contacts according to

616 predefined shear and tensile strength properties. A variant of the Voronoi code,

SC
617 Trigon, has recently been introduced where the rock slope is discretized into

618 triangular, not polygonal, blocks; as before, new stress-induced cracks are

U
619 developed when the intact rock properties on the Trigon elements are exceeded.

620 AN
Figure 18b shows the use of Trigon in the 2D UDEC code to simulate brittle fracture.

621 A DFN can be incorporated within the Voronoi or Trigon meshes to allow simulation
M
622 of rock bridge failure between pre-existing joints. A 3D Voronoi/Trigon method has

623 been developed by Gao and Stead (2014) to allow simulation of 3D fracture in rock
D

624 slopes, but has yet to be used for detailed rock slope analyses. The third method of
TE

625 simulating brittle fracture in rock slopes is through the use of PFC either in 2D or 3D.

626 The rock slope is simulated using particles that are bonded together; stress-induced
EP

627 failure of the particle bonds simulates the development of cracks in the rock. Wang

628 et al. (2003) describe the successful use of PFC (Itasca, 2014) to model footwall
C

629 slopes (Figure 18c), while Lorig et al. (2009) show the use of PFC with an
AC

630 incorporated DFN to model large rock slopes (Figure 18d). The combination of a

631 PFC with an incorporated DFN is referred to as a synthetic rock mass (SRM)

632 (Sainsbury et al. 2008). In this method, 3D SRM models are used to define the

633 strength of jointed rock masses up to 100 m in size, and, based on these strengths,

634 continuum 3D codes are used to model large open pit slopes up to 1 km in height.
ACCEPTED MANUSCRIPT

635 Finally, a recent development has been the lattice spring code approach, Slope

636 Model (Itasca, 2014). In this approach the particles and bonds within a PFC code are

637 replaced by nodes and springs. The method has been used both in the simulation of

638 natural rock slopes, such as the Vajont Slide in Italy (Figure 19), and large open pit

PT
639 slopes incorporating DFNs. Havaej et al. (2014), for example, show the use of

640 damage ellipsoids and Flinn plots in the analysis of brittle fracture modelling data

RI
641 from large open pits and landslides.

SC
642 A wide range of sophisticated geomechanical models exist for both 2D and 3D

643 modelling of rock slopes; however, their successful application requires careful

U
644 consideration of structural geology and tectonics:

645 •
AN
Highly tectonized weak rock mass may be amenable to continuum modelling,

646 whereas major discrete faults may dictate the use of discontinuum codes.
M
647 Highly foliated rocks may necessitate the use of special strength models
D

648 allowing for directional planes of weakness. Realistic simulation of the


TE

649 kinematics and selection of the appropriate model, as well as the choice of

650 peak and residual shear strengths and the directional shear strength
EP

651 properties, are intimately related to structure and tectonics.

652 • Geological structures may control not only the basal sliding surface but also
C

653 the lateral and rear release surfaces.


AC

654 • Rock mass strength and kinematics will vary and models must consider

655 structural domains.

656 • Discrete fracture networks will dictate the use of DFN model generators that

657 are able to simulate the true variation of fracture intensity and orientation

658 associated with faulting, folding, etc.


ACCEPTED MANUSCRIPT

659 • Groundwater pressures may be controlled by structures, which may result in

660 multiple groundwater tables or partitioning of the groundwater flow.

661 • The presence of faults or large-scale structures may have important controls

662 on slope failures in seismically active areas. Faults may act as sources of

PT
663 dynamic disturbance leading to failure; major discrete structures, if present in

664 the slope, may also affect amplification of seismic waves.

RI
665 5. Conclusions

SC
666 There are many similarities between structural geology and geotechnical

667 investigations. Both disciplines utilise field characterisation, geophysical, and remote

U
668 sensing methods to analyse features and investigate kinematics and mechanics. The

669
AN
aims of each discipline differ, however, one focussing on the history of the Earth’s

670 deformation and the other concerning soil and rock materials and how they interact
M
671 with the engineered world. Nonetheless, structural geology concepts and principles
D

672 are extremely valuable in geotechnical studies. Currently, common practice in slope
TE

673 design often neglects the investigation of the evolution of structural conditions

674 throughout a project’s lifespan. We advocate continued reassessment of structural


EP

675 features.

676 The inheritance of structural settings and features controls slope stability and
C

677 failure behaviour from the micro-scale to macro-scale (Figure 20). Structure can and
AC

678 often does affect fundamental rock mass properties, such as apparent cohesion,

679 frictional resistance, and shear strength, as well as failure kinematics and

680 mechanism. Small features such as centimetre-scale sedimentary structures,

681 fractures, and folds affect local behaviour of rock masses. Large features such as

682 decametre-scale folds and faults influence the geometry of rock failures, as well as
ACCEPTED MANUSCRIPT

683 their kinematics. In general, the orientations and dimensions of structures determine

684 whether an instability is a planar, wedge, toppling, rotational, or compound failure.

685 The locations of planes of weakness, such as faults or fold hinges, which increase

686 rock mass damage and thus degrade rock mass strength, can control failure location

PT
687 and sliding surface geometry and also influence landslide runout. In addition,

688 structures commonly affect hydrogeological conditions, a crucial factor in slope

RI
689 stability, acting as either water conduits or aquitards.

SC
690 Whether it is from the consideration of stress and strain techniques or the

691 delineation of major structures and their influence on the rock mass, structural

U
692 geology is of fundamental importance in rock slope investigations. Numerous studies

693 AN
have shown the role of major structures in defining the limits of landslides.

694 Consideration of mechanical stratigraphy, as discussed in structural geology, is


M
695 equally relevant to slope engineers when considering rock mass deformation and

696 strength variations in high mountain slopes. We believe that integrated structural
D

697 geological – geotechnical studies are essential to improve knowledge of landslide


TE

698 mechanisms. Recent landslide studies have shown the immense value of multi-

699 disciplinary studies involving rock engineers, engineering geologists, and structural
EP

700 geologists in both natural and engineered slopes.


C

701 References
AC

702 Alghalandis, Y.F., Dowd, P.A., Xu, C., 2014. Connectivity field: A measure for
703 characterising fracture networks. Mathematical Geoscience, DOI:
704 10.1007/s11004-014-9520-7.
705 Agliardi, F., Crosta, G.B., Frattini, P., Malusà, M.G., 2013. Giant non-catastrophic
706 landslides and the long-term exhumation of the European Alps. Earth and
707 Planetary Science Letters 365, 263-274.
708 Ambrosi, C., Crosta, G.B., 2011. Valley slope influence on deformation mechanisms
709 of rock slopes. In: Jaboyedoff, M. (Ed.), Slope Tectonics. Geological Society of
710 London, London, 215-233.
ACCEPTED MANUSCRIPT

711 Baczynski, N.R.P., 2000. STEPSIM4 “step-path” method for slope risks. In:
712 Proceedings of the International Conference on Geotechnical and Geological
713 Engineering (GeoEng 2000), Melbourne, Australia, November 2000. Technomic
714 Publishing Co., Lancaster, 86-92.
715 Badger, T.C., 2002. Fracturing within anticlines and its kinematic control on slope
716 stability. Environmental and Engineering Geoscience 8, 19-33.
717 Barton, N.R., 1976. The shear strength of rock and rock joints. International Journal
718 of Rock Mechanics and Mining Sciences & Geomechanics Abstracts 13, 1-24.

PT
719 Barton, N.R., 2011. From empiricism, through theory, to problem solving in rock
720 engineering. In: Qian, Q., Zhou, Y. (Eds.), Harmonising Rock Engineering and
721 the Environment. Taylor & Francis, London, 3-14.

RI
722 Barton, N.R., Choubey, V., 1977. The shear strength of rock joints in theory and
723 practice. Rock Mechanics 10, 1-54.
724 Benko, B., 1997. Numerical modelling of complex slope deformations. Ph.D. thesis,

SC
725 University of Saskatchewan.
726 Benko, B., Stead, D., 1998. The Frank slide: A re-examination of the failure
727 mechanism. Canadian Geotechnical Journal 35, 299-311.
728 Blikra, L.H., 2012. The Åknes rockslide, Norway. In: Clague, J.J., Stead, D. (Eds.),

U
729 Landslides: Types, Mechanisms and Modelling. Cambridge University Press,
730 New York, 323-335.
731
732
733
AN
Brian, M.J., Rosser, N.J., Norman, E.C., Petley, D.N., 2014. Are microseismic
ground displacements a significant geomorphic agent? Geomorphology 207, 161-
173.
734 Brideau, M.-A., 2005. The influence of tectonic structures on rock mass quality and
M
735 implications for rock slope stability. Master’s thesis, Simon Fraser University.
736 Brideau, M.-A., 2010. Three-dimensional kinematic controls on rock slope stability
737 conditions. Ph.D. thesis, Simon Fraser University.
738 Brideau, M.-A., McDougall, S., Stead, D., Evans, S.G., Couture, R., Turner, K., 2012.
D

739 Three-dimensional Distinct Element modelling and dynamic modelling analysis of


740 a landslide in gneissic rock, British Columbia, Canada. Bulletin of Engineering
TE

741 Geology and the Environment 71, 467-486.


742 Brideau, M.-A., Stead, D., Couture, R., 2006. Structural and engineering geology of
743 the East Gate Landslide, Purcell Mountains, British Columbia, Canada.
744 Engineering Geology 84, 183-206.
EP

745 Brideau, M.-A., Yan, M., Stead, D., 2009. The role of tectonic damage and brittle
746 rock fracture in the development of large rock slope failures. Geomorphology
747 103, 30-49.
748 Bromhead, E.N., 1992. The Stability of Slopes. Blackie Academic & Professional,
C

749 Glasgow.
750 Cai, M., Kaiser, P.K., Uno, H., Tasaka, Y., Minami, M., 2004. Estimation of rock
AC

751 mass deformation modulus and strength of jointed hard rock masses using the
752 GSI system. International Journal of Rock Mechanics and Mining Sciences 41, 3-
753 19.
754 Cappa, F., Guglielmi, Y., Soukatchoff, V.M., Mudry, J., Bertrand, C., Charmoille, A.,
755 2004. Hydromechanical modelling of a large moving rock slope inferred from
756 slope levelling coupled to spring long-term hydrochemical monitoring: example of
757 the La Clapière landslide (Southern Alps, France). Journal of Hydrology 291, 67-
758 90.
ACCEPTED MANUSCRIPT

759 Clayton, M.A., 2014. Characterization and analysis of the Mitchell Creek landslide: A
760 large-scale rock slope instability in north-western British Columbia. Master’s
761 thesis, Simon Fraser University.
762 Cruden, D.M., Krahn, J., 1978. Frank Rockslide, Alberta, Canada. In: Voight, B.
763 (Ed.), Rockslides and Avalanches, Elsevier, Amsterdam, 97-112.
764 Damjanac, B., Varun, Lorig, L., 2013. Seismic stability of large open pit slopes and
765 pseudo-static analysis. In: Dight, P.M. (Ed.), Slope Stability 2013. Australian
766 Centre for Geomechanics, Perth, 1203-1216.

PT
767 Diederichs, M.S., 2003. Manuel Rocha medal recipient rock fracture and collapse
768 under low confinement conditions. Rock Mechanics and Rock Engineering 36,
769 339-381.

RI
770 Dowd, .A., Xu, C., Mardia, K.V., Fowell, R.J., 2007. A comparison of methods for the
771 stochastic simulation of rock fractures. Journal of Mathematical Geology 39, 697–
772 714.

SC
773 Duncan, J.M., Wright, S.G., 2005. Soil Strength and Slope Stability. John Wiley &
774 Sons, Inc., Hoboken, New Jersey.
775 Eberhardt, E., Stead, D., Coggan, J.S., 2004. Numerical analysis of initiation and
776 progressive failure in natural rock slopes – the 1991 Randa rockslide.

U
777 International Journal of Rock Mechanics and Mining 41, 69-87.
778 Fleuty, M.J., 1964. The description of folds. Proceedings of the Geologists’
779
780
781
Association 75, 461-492.
AN
Fossen, H., 2010. Structural Geology. Cambridge University Press, Cambridge.
Franz, J., 2009. An investigation of combined failure mechanisms in large scale open
782 pit slopes. Ph.D. thesis, University of New South Wales.
M
783 Ganerød, G.V., Grøneng, G., Rønning, J.S., Dalsegg, E., Elvebakk, H., Tønnesen,
784 J.F., Kveldsvik, V., Eiken, T., Blikra, L.H., Braathen, A., 2008. Geological model
785 of the Åknes rockslide, western Norway. Engineering Geology 102, 1-18.
786 Gao, F.Q., Stead, D., 2014. The application of a modified Voronoi logic to brittle
D

787 fracture modelling at the laboratory and field scale. International Journal of Rock
788 Mechanics and Mining Sciences 68, 1-14.
TE

789 Gischig, V., Löw, S., Kos, A., Moore, J.R., Raetzo, H., Lemy, F., 2009. Identification
790 of active release planes using ground-based differential InSAR at the Randa rock
791 slope instability, Switzerland. Natural Hazards and Earth System Sciences 9,
792 2027-2038.
EP

793 Gischig, V., Amann, F., Moore, J.R., Löw, S., Eisenbeiss, H., Stempfhuber, W.,
794 2011. Composite rock slope kinematics at the current Randa instability,
795 Switzerland, based on remote sensing and numerical modelling. Engineering
796 Geology 118, 37-53.
C

797 Glastonbury, J., Fell, R., 2000. Report on the analysis of “rapid” natural rock slope
798 failures. University of New South Wales, School of Civil and Environmental
AC

799 Engineering Report No. R390.


800 Golder, 2012. Fracman (v. 7.4). Golder Associates, Redmond, WA, USA.
801 Goodman, R.E., 2013. Toppling – A fundamental failure mode in discontinuous
802 materials, Description and analysis. Geo Congress, 2338-2368.
803 Grøneng, G., Christiansen, H.H., Nilsen, B., Blikra, L.H., 2011. Meteorological effects
804 on seasonal displacements of the Åknes rockslide, western Norway. Landslides
805 8, 1-15.
806 Grøneng, G., Lu, M., Nilsen, B., Jenssen, A.K., 2010. Modelling of time-dependent
807 behaviour of the basal sliding surface of the Åknes rockslide area in western
808 Norway. Engineering Geology 114, 412-422.
ACCEPTED MANUSCRIPT

809 Hamdi, P., Stead, D., Elmo, D., 2013. Numerical simulation of damage during
810 laboratory testing on rock using a 3D-FEM/DEM approach. In: Proceedings
811 ARMA 2013, ARMA Paper 13-583.
812 Hammah, R.E., Yacoub, T., Curran, J.H., 2009. Variation of failure mechanisms of
813 slopes in jointed rock masses with changing scale. In: Diederichs, M., Grasselli,
814 G. (Eds.), Proceedings of the 3rd CANUS Rock Mechanics Symposium, Toronto,
815 ON. Canadian Association of Rock Mechanics, Paper 3956.
816 Haneberg, W.C., 2007. Directional roughness profiles from three-dimensional

PT
817 photogrammetric or laser scanner point clouds. In: Eberhardt, E., Stead, D.,
818 Morrison, T. (Eds.), Rock Mechanics: Meeting Society’s Challenges and
819 Demands. Proceedings of the 1st Canada-US Rock Mechanics Symposium,

RI
820 Vancouver, BC, Canada, 101-106.
821 Havaej, M., Stead., D., Mayer, J., Wolter, A., 2014. Modelling the relationship
822 between failure kinematics and slope damage in high rock slopes using a lattice

SC
823 scheme approach. ARMA 2014.
824 Havaej, M., Wolter, A., Stead, D., Tuckey, Z., Lorig, L., Eberhardt, E., 2013.
825 Incorporating brittle fracture in three-dimensional modelling of high rock slopes.
826 In: Dight, P.M. (Ed.), Slope Stability 2013, Brisbane, September, 2013. Australian

U
827 Centre for Geomechanics, Perth.
828 Hencher, S.R., 2013. Characterizing discontinuities in naturally fractured outcrop
829
830
831
AN
analogues and rock core: The need to consider fracture development over
geological time. In: Spence, G.H., Redfern, J., Aguilera, R., Bevan, T.G.,
Cosgrove, J.W., Couples, G.D., Daniel, J.-M. (Eds.), Geological Society, London,
832 Special Publications 374, 113-123.
M
833 Hencher, S.R., Richards, L.R., 1982. The basic frictional resistance of sheeting joints
834 in Hong Kong Granite. Hong Kong Engineering 11, 21-25.
835 Hencher, S.R., Richards, L.R., 2014. Assessing the shear strength of rock
836 discontinuities at laboratory and field scales. Rock Mechanics and Rock
D

837 Engineering. DOI: 10.1007/s00603-014-0633-6.


838 Hencher, S.R., Lee, S.G., Carter, T.G., Richards, L.R., 2011. Sheeting joints:
TE

839 Characterisation, shear strength and engineering. Rock Mechanics and Rock
840 Engineering 44, 1-22.
841 Hoek, E., Bray, J. W., 1981. Rock slope engineering. Taylor & Francis.
842 Hoek, E., Brown, E.T., 1980. Empirical strength criterion for rock masses. Journal of
EP

843 Geotechnical Engineering Division ASCE, 1013-1025.


844 Hoek, E., Carter, T.G., Diederichs, M.S., 2013. Quantification of the Geological
845 Strength Index Chart. In: Proceedings of the 47th US Rock
846 Mechanics/Geomechanics Symposium, San Francisco, CA, USA, June 23-26,
C

847 2013. ARMA, Paper 13-672.


848 Hoek, E., Hutchinson, J., Kalenchuk, K., Diederichs, M., 2009. Appendix 3: Influence
AC

849 of in situ stresses on open pit design. In: Read, J., Stacey, P. (Eds.), Guidelines
850 for Open Pit Slope Design. CSIRO Publishing, Collingwood, Australia, 437-445.
851 Hoek, E., Kaiser, P.K., Bawden, W.F., 1995. Support of Underground Excavations in
852 Hard Rock. A.A. Balkema, Rotterdam, Netherlands.
853 Humair, F., Pedrazzini, A., Epard, J.-L., Froese, C.R., Jaboyedoff, M., 2013.
854 Structural characterization of Turtle Mountain anticline (Alberta, Canada) and
855 impact on rock slope failure. Tectonophysics 605, 133-148.
856 Hungr, O., Amann, F., 2011. Limit equilibrium of asymmetric laterally constrained
857 rockslides. International Journal of Rock Mechanics and Mining Sciences 48,
858 748-758.
ACCEPTED MANUSCRIPT

859 Itasca, 2014. 3DEC (v. 5.0), FLAC2D (v. 7.0), FLAC3D (v. 5.0), PFC (v. 5.0), Slope
860 Model (v. 2.9), UDEC (v. 6.0). Itasca Consulting Group. Minneapolis, MN, USA.
861 Jaboyedoff, M., Metzger, R., Oppikofer, T., Couture, R., Derron, M.-H., Locat, J.,
862 Turmel, D., 2007. New insight techniques to analyze rock-slope relief using DEM
863 and 3D-imaging cloud points: COLTOP-3D software. In: Eberhardt, E., Stead, D.,
864 and Morrison, T. (Eds.) Rock Mechanics: Meeting Society’s Challenges and
865 Demands. Taylor & Francis, London, 61-68.
866 Jennings, J.E.B., 1970. A mathematical theory for the calculation of the stability of

PT
867 slopes in open cast mines. In: Proceedings of the Symposium on the Planning of
868 Open Pit Mines, Johannesburg, South Africa, August/September 1970. Balkema,
869 Cape Town, 87–102.

RI
870 Kalenchuk, K.S., 2010. Multi-dimensional analysis of large, complex slope instability.
871 Ph.D. thesis, Queen’s University.
872 Kinakin, D., Stead, D., 2005. Analysis of the distributions of stress in natural ridge

SC
873 forms: Implications for the deformation mechanisms of rock slopes and the
874 formation of sackung. Geomorphology 65, 85-100.
875 Lato, M., Diederichs, M.S., Hutchinson, D.J., Harrap, R., 2009. Optimization of
876 LiDAR scanning and processing for automated structural evaluation of

U
877 discontinuities in rock masses. International Journal of Rock Mechanics and
878 Mining, 46, 194-199.
879
880
881
AN
Lebourg, T., Mickael, H., Hervé, J., Samyr, E.B.B., Thomas, B., Swann, Z.,
Emmanuel, T., Maurin Jr., V., 2010. Temporal evolution of weathered cataclastic
material in gravitational faults of the La Clapiere deep-seated landslide by
882 mechanical approach. Landslides 8, 241-252.
M
883 Leith, K., Moore, J.R., Amann, F., Löw, S., 2014a. Subglacial extensional fracture
884 development and implications for Alpine Valley evolution. Journal of Geophysical
885 Research: Earth Surface 119, 62-81.
886 Leith, K., Moore, J.R., Amann, F., Löw, S., 2014b. In situ stress control on
D

887 microcrack generation and macroscopic extensional fracture in exhuming


888 bedrock. Journal of Geophysical Research: Solid Earth 119, 594-615.
TE

889 Lorig, L., Stacey, P., Read, J., 2009. Slope design methods. In: Read, J., Stacey, P.
890 (Eds.), Guidelines for Open Pit Slope Design. CSIRO Publishing, Collingwood,
891 237-264.
892 Löw, S., Gischig, V., Willenberg, H., Alpiger, A., Moore, J.R., 2012. Randa:
EP

893 Kinematics and driving mechanisms of a large complex rockslide. In: Clague,
894 J.J., Stead, D. (Eds.), Landslides: Types, Mechanisms and Modelling. Cambridge
895 University Press, New York, 297-309.
896 Marinos, P., Hoek, E., 2000. GSI: A geologically friendly tool for rock mass strength
C

897 estimation. In: Proceedings of the GeoEng2000 at the International Conference


898 on Geotechnical and Geological Engineering, Melbourne. Technomic Publishers,
AC

899 Lancaster, UK, 1422-1446.


900 Massironi, M., Zampieri, D., Superchi, L., Bistacchi, A., Ravagnan, R., Bergamo, A.,
901 Ghirotti, M., Genevois, R., 2013. Geological structures of the Vajont landslide.
902 Italian Journal of Engineering Geology and Environment - Book Series 6, 573-
903 582.
904 Midland Valley, 2014. MOVE4D. Midland Valley Exploration Ltd.
905 Milmo, P., Rogers, S., Raynor, M., Beale, G., 2014. Appendix 3: Case study: Diavik
906 North-west wall. In: Read, J., Beale, G. (Eds.), Guidelines for Evaluating Water in
907 Pit Slope Stability. CSIRO Publishing, Collingwood, 436-485.
ACCEPTED MANUSCRIPT

908 Moore, J.R., Gischig, V., Burjanek, J., Löw, S., Fäh, D., 2011. Site effects in unstable
909 rock slopes: Dynamic behaviour of the Randa instability (Switzerland). Bulletin of
910 the Seismological Society of America 101, 3110-3116.
911 Newman, S., 2013. Deep-seated gravitational slope deformations near the Trans-
912 Alaska Pipeline, east-central Alaska Range. Master’s thesis, Simon Fraser
913 University.
914 Nicksiar, M., Martin, C.D., 2014. Factors affecting crack initiation in low porosity
915 crystalline rocks. Rock Mechanics and Rock Engineering 47, 1165-1181.

PT
916 Pedrazzini, A., Jaboyedoff, M., Froese, C.R., Langenberg, C.W., Moreno, F., 2011.
917 Structural analysis of Turtle Mountain: Origin and influence of fractures in the
918 development of rock slope failures. Geological Society, London, Special

RI
919 Publications 351, 163-183.
920 Petley, D., 2012. Remote sensing techniques and landslides. In: Clague, J.J., Stead,
921 D. (Eds.), Landslides: Types, Mechanisms and Modelling. Cambridge University

SC
922 Press, New York, 159-171.
923 Priest, S.D., 1980. The use of inclined hemisphere projection methods for the
924 determination of kinematic feasibility, slide direction and volume of rock blocks.
925 International Journal of Rock Mechanics and Mining Sciences & Geomechanics

U
926 Abstracts 17, 1-23.
927 Ramsay, J.G., 1967. Fold and fracturing of rocks. McGraw-Hill, New York.
928
929
930
Publishing, Collingwood.
AN
Read, J., Stacey, P. (Eds.), 2009. Guidelines for Open Pit Slope Design. CSIRO

Richards, L.R., Leg., G.M.M., Whittle, R.A., 1978. Appraisal of stability conditions in
931 rock slopes. In: Bell, F.G. (Ed.), Foundation Engineering in Difficult Ground,
M
932 Butterworths, London, 449-512.
933 Rocscience, 2014. DIPS (v. 6.0), Phase2 (v. 9.0), RS3 (v. 1.0), Swedge (v. 6.0).
934 Rocscience Inc., Toronto, ON, Canada.
935 Sainsbury, B., Pierce, M.E., Mas Ivars, D., 2008. Analysis of caving behaviour using
D

936 a synthetic rock mass-ubiquitous joint rock mass modelling technique. In:
937 Proceedings of the 1st Southern Hemisphere International Rock Mechanics
TE

938 Symposium (SHIRMS 2008), Perth, Australia. International Society for Rock
939 Mechanics, 243-253.
940 Shipton, Z.K., Cowie, P.A., 2003. A conceptual model for the origin of fault damage
941 zone structures in high-porosity sandstone. Journal of Structural Geology 25,
EP

942 333-344.
943 Shipton, Z.K., Cowie, P.A., 2008. Damage zone and slip-surface evolution over µm
944 to km scales in high-porosity Navajo sandstone, Utah. Journal of Structural
945 Geology 23, 1825-1844.
C

946 Stead, D., Coggan, J.S., 2012. Numerical modelling of rock-slope instability. In:
947 Clague, J.J., Stead, D. (Eds.), Landslides: Types, Mechanisms and Modelling.
AC

948 Cambridge University Press, New York, 144-158.


949 Stead, D., Eberhardt, E. 2013. Understanding the mechanics of large landslides.
950 Invited keynote and paper. Italian Journal of Engineering Geology and
951 Environment - Book Series 6, 85-112.
952 Stead, D., Eberhardt, E., Coggan, J.S., 2006. Developments in the characterization
953 of complex rock slope deformation and failure using numerical modelling
954 techniques. Engineering Geology 83, 217-235.
955 Sturzenegger, M., Stead, D., 2009. Quantifying discontinuity orientation and
956 persistence on high mountain rock slopes and large landslides using terrestrial
ACCEPTED MANUSCRIPT

957 remote sensing techniques. Natural Hazards and Earth System Sciences 9, 267-
958 287.
959 Styles, T.D., 2009. Numerical modelling and analysis of slope stability within fracture
960 dominated rock masses. Ph.D. thesis, Camborne School of Mines, University of
961 Exeter.
962 Tatone, B.S.A., Grasselli, G., 2010. A new 2D discontinuity roughness parameter
963 and its correlation with JRC. International Journal of Rock Mechanics and Mining
964 Sciences 47, 1391-1400.

PT
965 Thiessen, R., Means, W.D., 1980. Classification of fold interference patterns: A re-
966 examination. Journal of Structural Geology 2, 311-326.
967 Tuckey, Z., 2012. An integrated field mapping-numerical modelling approach to

RI
968 characterising discontinuity persistence and intact rock bridges in large open pit
969 slopes. Master’s thesis, Simon Fraser University.
970 Tuckey, Z., Stead, D., Eberhardt, E., 2013. An integrated approach for

SC
971 understanding uncertainty of discontinuity persistence and intact rock bridges in
972 large open pit slopes, Slope Stability 2013, Brisbane, Sept 2013.
973 Vengeon, J.-M., Giraud, A., Antoine, P., Rochet, L., 1999. Contribution à l’analyse de
974 la deformation et de la rupture des grands versants rocheux en terrain

U
975 cristallophyllien. Canadian Geotechnical Journal 36, 1123-1136.
976 Vivas Becerra, J., 2014. Groundwater characterization and modelling in natural and
977
978
979
AN
open pit rock slopes. Master’s thesis, Simon Fraser University.
Vyazmensky, A., Stead, D., Elmo, D., Moss, A., 2010. Numerical analysis of block
caving-induced instability in large open pit slopes: A finite element/discrete
980 element approach. Rock Mechanics and Rock Engineering 43, 21-39.
M
981 Wang, C., Tannant, D.D., Lilly, P.A., 2003. Numerical analysis of the stability of
982 heavily jointed rock slopes using PFC2D. International Journal of Rock
983 Mechanics and Mining Sciences 3, 415-424.
984 Willenberg, H., Löw, S., Eberhardt, E., Evans, K.F., Spillman, T., Heinke, B., Maurer,
D

985 H., Green, A.G., 2008. Internal structure and deformation of an unstable
986 crystalline rock mass above Randa (Switzerland): Part I – internal structure from
TE

987 integrated geological and geophysical investigations. Engineering Geology 101,


988 1-14.
989 Wolter, A., 2014. Characterisation of large catastrophic landslides using an
990 integrated field, remote sensing, and numerical modelling approach. Ph.D. thesis,
EP

991 Simon Fraser University.


992 Wolter, A., Stead, D., Clague, J.J., 2014. A morphologic characterisation of the 1963
993 Vajont Slide, Italy, using long-range terrestrial photogrammetry. Geomorphology
994 206, 147-164.
C

995 Wyllie, D.C., Mah, C.W., 2004. Rock Slope Engineering: Civil and Mining (4th Ed.).
996 Spon Press, New York.
AC

997 Xu, C., Dowd, P., 2010. A new computer code for discrete fracture network
998 modelling. Computers and Geosciences 36, 292-301.
999 Yan, M., 2008. Numerical modelling of brittle fracture and step-path failure: from
1000 laboratory to rock slope scale. Ph.D. thesis, Simon Fraser University.
1001 Ziegler, M., Löw, S., Bahat, D., 2014. Growth of exfoliation joints and near-surface
1002 stress orientations inferred from fractographic markings observed in the upper
1003 Aar valley (Swiss Alps). Tectonophysics 626, 1-20.
1004 Ziegler, M., Löw, S., Moore, J.R., 2013. Distribution and inferred age of exfoliation
1005 joints in the Aar Granite of the central Swiss Alps and relationship to Quaternary
1006 landscape evolution. Geomorphology 201, 344-362.
ACCEPTED MANUSCRIPT

1007 Zoback, M.L., 1992. First- and second-order patterns of stress in the lithosphere:
1008 The World Stress Map Project. Journal of Geophysical Research: Solid Earth 97,
1009 11703-11728.
1010

PT
RI
U SC
AN
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

1011 Figure Captions

1012 Figure 1. Key rock failure modes considered in slope stability analysis: a)
1013 planar/translational sliding, b) toppling, and c) wedge sliding. d) Multi-planar
1014 translational failure (Palliser rockslide), demonstrating the complexity slope
1015 engineers may encounter.

PT
1016 Figure 2. Classification of failure modes related to structures. a) Large rock glide, b)
1017 rough translational slide, c) planar translational slide, d) toe buckling translational
1018 slide, e) biplanar compound slide, f) curved compound slide, g) toppling failure, h)
1019 irregular compound slide (modified after Glastonbury and Fell, 2000).

RI
1020 Figure 3. Open pit slope showing influence of structural geology, with translational
1021 failure mechanism and low-angle thrust faulting influencing the footwall instability

SC
1022 (along bedding). Note the seepage at the fold hinge.
1023 Figure 4. Modification of Fleuty’s (1964) fold classification diagram to indicate slope
1024 stability and failure complexity. Kinematic complexity is highest in the centre of the

U
1025 diagram, where multiple modes of failure may be combined. Inset shows an example
1026 of the interaction between slopes and tectonic structure at Mount Kidd, Canada.
1027
1028
AN
Figure 5. Modification of fold tightness classification, showing how damage increases
with increasing fold tightness (after Fossen, 2010). Transverse, longitudinal, and
1029 converse discontinuities may be present in open folds, whereas tight folds may
1030 develop foliation and cleavage.
M
1031 Figure 6. Scales of roughness at the Vajont Slide, ranging from a) global (macro)
1032 scale synclines (MS = Massalezza Syncline with hinge trending N, ES = Erto
D

1033 Syncline trending E) to meso-scale interference patterns (b), parasitic folds (c), and
1034 micro-scale concretions (d). CT = Col Tramontin Fault, and CE = Col delle Erghene
1035 Fault in a).
TE

1036 Figure 7. Fold and slope interaction creating slope instability (indicated by grey
1037 areas). S0 = bedding, S1 = discontinuity set 1, S2 = discontinuity set 2. a) Bedding
1038 dips parallel to slope and orthogonal discontinuity set acts as rear release, b)
EP

1039 bedding dips into slope and orthogonal discontinuity set acts as main sliding surface,
1040 and c) slope face intersects with fold hinge and bedding and orthogonal discontinuity
1041 set create a biplanar failure surface (modified from Badger, 2002).
C

1042 Figure 8. The 1903 Frank Slide in Alberta, Canada. The South Peak area is still
1043 unstable and subject to ongoing monitoring.
AC

1044 Figure 9. a) Example of drag folding-induced slope failure at a former UK coal mine.
1045 Fault outcrops in the trees at the back of the failure, and dips parallel to general
1046 sliding direction indicated by the white arrow. Note steepening of bedding adjacent to
1047 fault in b), which resulted in additional driving forces for failure and ultimately slope
1048 instability.
1049 Figure 10. The 1991 Randa rock slope failure in Switzerland. Foliation planes dip
1050 into the slope (Gischig et al., 2011).
ACCEPTED MANUSCRIPT

1051 Figure 11. La Clapière landslide, France (inset), an active part of a larger DSGSD
1052 complex (from Lebourg et al., 2010).
1053 Figure 12. Kinematic analysis conducted in the stereographic software DIPS
1054 (Rocscience, 2014), showing daylight envelope containing poles to planar failures.
1055 Friction cone delimits joints with dip greater than assumed friction angle, and hence
1056 potential sliding.

PT
1057 Figure 13. Two pentahedral wedge geometries in Swedge (Rocscience, 2014).
1058 Figure 14. Geological Strength Index (GSI) chart, based on block size and joint
1059 condition (from Hoek et al., 1995).

RI
1060 Figure 15. GSI chart for heterogeneous rock masses such as flysch (from Marinos
1061 and Hoek, 2000).

SC
1062 Figure 16. Examples of continuum analysis. a) Phase2 simulations of a 60-m-high
1063 slope using different joint networks – continuous (left), non-persistent (centre), and
1064 Voronoi (right) (modified after Hammah et al., 2009). b) RS3 model of an open pit
1065 (from Rocscience website, http://www.rocscience.com). c) FLAC2D model of the

U
1066 Frank Slide showing deformation along a presumed failure surface (after Benko and
1067 Stead, 1998). d) FLAC3D model geometry of the Åknes, Norway, rock slope
1068
1069
AN
instability (after Grøneng et al., 2010).
Figure 17. Examples of discontinuum analysis. a) UDEC model of the Vajont Slide,
1070 Italy (after Wolter, 2014). Note that the sliding surface follows folded bedding. b)
M
1071 3DEC model of the Beauregard, Italy, landslide, showing discrete movement blocks
1072 (after Kalenchuk, 2010). c) UDEC model of the Frank Slide, Canada (after Benko
1073 and Stead, 1998).
D

1074 Figure 18. Examples of brittle fracture and damage modelling. a) Progressive
1075 development of failure due to excavation of an open pit in an FDEM simulation (after
TE

1076 Vyazmensky et al., 2010). b) Brittle fracture within a Trigon model of the 1963 Vajont
1077 Slide, Italy, in UDEC, where red lines indicate tensile failure, and green lines indicate
1078 shear failure (after Wolter, 2014). c) Failure of a slope with 70% fracture persistence
1079 in PFC2D (after Wang et al., 2003). d) Combination of particle flow code and DFN in
EP

1080 an open pit slope (after Lorig et al., 2009).


1081 Figure 19. The 1963 Vajont Slide, Italy, simulated using an imported DEM and
1082 assumed sliding surface in the new lattice-spring code Slope Model (after Wolter,
C

1083 2014; Havaej et al., 2013).


AC

1084 Figure 20. Summary of the influence of structure on rock slope stability and failure
1085 mechanisms.
1086 Table Captions
1087 Table 1. Summary of case studies presented in the text.
1088 Table 2. Summary of numerical modelling approaches used to simulate brittle
1089 fracture.
ACCEPTED MANUSCRIPT

Table 1
Name Dominant Type Volume (m3) Lithology Structural Control Mechanisms

PT
Åknes, Norway DSGSD 35,000,000 gneiss foliation, 6 DS, faults, folds sliding
> 1960 - present rockslide biotite schist rear release - tension crack intact rock fracture
gouge lateral release - faults creep

RI
Beauregard, Italy DSGSD 663,000,000 mica schist sub-vertical DS sliding
1951 - present paragneiss release - fractures, trough toppling

SC
Downie, British Columbia DSGSD 1,500,000,000 micaschist 4 main DS, foliation, folds, fault creep
10 000 yr ago gneiss rear release - ridge crest sliding

U
quartzite lateral release - joint sets
East Gate, British Columbia rockslide n/a phyllite 4 main DS, foliation, cleavage sliding

AN
1997 debris flows release - joint sets flow

Eiger, Switzerland rock collapse 2,000,000 limestone 3 main DS sliding

M
Jul. 13, 2006 release - joint sets rockfall

D
Flims, Switzerland rockslide 8.5- limestone faults, folds, foliation sliding
8500 yr ago 12,000,000,000 release - joint sets, faults toppling
TE
Frank, Alberta rock avalanche 30,000,000 limestone 3 DS, anticline, faults sliding
Apr. 29, 1903 clastics rear release - fold hinge toppling
EP

lateral release - joint sets wedge failure


progressive deformation subsidence
C

Goldau, Switzerland rockslide 36,000,000 marl joints, bedding sliding


Sept. 2, 1806 conglomerate release - joint sets creep
AC

Hope, British Columbia rockslide 47,000,000 greenstone foliation, DS, shear zones, faults sliding
1965 felsite rear release - fault flow
lateral release - joint sets, faults
ACCEPTED MANUSCRIPT

Name Dominant Type Volume (m3) Lithology Structural Control Mechanisms


Köfels, Austria rock avalanche 2,100,000,000 para/orthogneiss fault, DS, foliation sliding

PT
8700 yr ago schist release - foliation, joint sets flow

La Clapière, France creep 60,000,000 gneiss foliation, fold, faults toppling

RI
1900 - present DSGSD schist fold alters foliation orientation sliding
diorite rear release - fractures creep

SC
lateral release - fractures, fold
Madison, Montana rockslide 20,000,000 gneiss 3-4 DS, foliation, regional faults composite sliding
Aug. 17, 1959 schist rear release - ridge crest wedge failure

U
dolostone lateral release - gullies (creep)
earthquake

AN
Mt Steele, Yukon rock/ice avalanche 27-80,000,000 granodiorite fault, 3 main DS sliding
Jul. 24, 2007 (5/95%) diorite release - joint sets flow
gabbro earthquake

M
Palliser, Alberta rockslide 8,000,000 cherty carbonates 5 DS, regional thrust, syncline sliding
prehistoric (2 events) 20,000,000 syncline curves bedding as base wedge failure

D
release - joint sets
Randa, Switzerland rockslide (2-stage) 30,000,000 ortho/paragneiss
TE 3 main DS, fault, schistosity sliding
Apr. 18, 1991 DSGSD rear release - joint sets toppling
May 9, 1991 3-9,000,000 lateral release - joint sets, fault
present
EP

Vajont, Italy rockslide 270,000,000 carbonates 9 DS, folds, faults sliding, rotation
Oct. 9, 1963 clay rear release - tension crack, fault wedge failure
C

lateral release - fault (E only) creep


Val Pola, Italy rock avalanche 34,000,000 diorite, gabbro 2 main DS, fault wedge failure
AC

Jul. 28, 1987 quartzite rear release - joint sets, faults sliding
gneiss, lateral release - fault, joint sets
amphibolite
ACCEPTED MANUSCRIPT

1090 Table 2
Approach Advantages Limitations Comments
Finite-Discrete • simulates intact rock fracture in rock • slope modelling mainly Successfully used in
Element Hybrid slopes from continuum to discontinuum confined to 2D to-date with modelling both natural
(Elfen/YGEO) • possible to fracture across or around future development in 3D and engineering rock
triangular elements pending slopes (Eberhardt et
• possible to include influence of • simulation of groundwater al., 2004; Stead et al.
groundwater table (Styles 2009) pressure dissipation in slopes 2006).

PT
• DFNs can be incorporated within rock varies in different FDEM
slope models models
• directional weakness may be simulated
Distinct Element • simulates intact rock fracture associated • careful calibration required of Used successfully in

RI
+ Voronoi/Trigon with slope failure, can incorporate Voronoi polygon/Trigon contact modelling of natural
(UDEC 2D, 3DEC polygons within bedded/through-going properties rock slopes and
3D) jointed rock masses • polygon size may influence engineered rock
• possible to incorporate DFNs, but adds kinematics of movement slopes. Further

SC
complexity • may need to edit DFNs to research required on
• triangular elements allow increased allow solution calibration of model
kinematic freedom • limited experience in use of properties.
• possible to incorporate influence of Trigon modelling

U
groundwater pressure in UDEC Voronoi • long model run times for large-
and Trigon simulation of slopes (Vivas scale problems especially for
2014) 3D problems which may not be

Particle Flow
Codes
AN
• brittle fracture of slopes can be modelled
through breaking of bonds between
feasible in some cases
• require careful calibration
• large rock slope models
Used to successfully
model footwall failures
(PFC2D/PFC3D) particles in 2D or spheres in 3D possible in 2D but in mines, large rock
M
• ability to include DFNs and groundwater computationally expensive in slopes and subsequent
pressures 3D runout and large open
• SRM testing possible • limited use in 3D models to pits slope.
date
D

SRM approach • derivation of large scale rock mass • SRM testing programs Successful use
(PFC3D/DFN/ properties through simulated brittle required for important predominantly for major
FLAC3D) fracture testing of large scale uniaxial, lithological units within rock mine slopes to date.
TE

triaxial and tensile tests using PFC3D + slopes


DFN (SRM) • fracture does not occur in final
• properties used for subsequent continuum model of rock slope
modelling in continuum 3D models
EP

(FLAC3D + directional strain softening)


of large-scale (800 m high) rock slopes
Lattice Spring • models brittle fracture through breaking • limited to small-scale New code with
Models of springs between nodes displacements compared to significant potential in
(SlopeModel) • allows simulation of large-scale rock UDEC/3DEC future rock slope
C

slopes in 3D • typical model properties modelling. Successfully


• ability to import DFNs provided but calibration used in modelling of
AC

• possible to undertake coupled hydro- required large open pit slopes


mechanical-brittle fracture modelling of a • runtime for hydro-mechanical and high natural
rock slope models may limit size of landslides.
problem
1091
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
EP
C
AC

Você também pode gostar