You are on page 1of 344

Numerical Modelling and Analysis of

slope stability within fracture dominated


rock masses

Thomas Daniel Styles BSc. (Hons) ACSM

Submitted by Thomas Styles to the University of Exeter as a thesis for the degree of
Doctor of Philosophy in Earth Resources in the School of Geography, Archaeology and
Earth Resources, January 2009.

This thesis is available for Library use on the understanding that it is copyright material
and that no quotation from the thesis may be published without prior
acknowledgement.

I certify that all material in this thesis, which is not my own work has been identified and
that no material is included for which, a degree has previously been conferred upon
me.
...................................... Thomas Styles

i
Abstract
Numerical modelling of rock slopes can involve a number and variety of techniques, the
selection and requirement of which depends on the factors deemed to control the
potential for instability. This thesis presents a number of case studies involving slopes
in fractured rock, encompassing a range of scales. The case study slopes have
provided a means to question the way in which particular slope instabilities should be
analysed. Currently there are few methods available for analysing the complex
behaviour within slopes of fractured rock. A review of available techniques is given
within this thesis, with the use of limit equilibrium, finite element and hybrid methods, to
highlight their specific advantages and limitations for the chosen case study slopes.

By modelling slope instability within fractured rock, the understanding of both discrete
and mass behaviour increases considerably. Numerical modelling can therefore be
used as a tool to help improve both the safety and efficiency of open pit mining and the
management of natural rock slopes.

The emphasis of the numerical modelling used in this thesis, was to assess the ability
of a particular comprehensive dynamic-based code, ELFEN, for modelling fractured
rock slopes. In addition, a principal objective of the research was to test the newly
developed groundwater version of the code. Investigations revealed ELFEN to be
effective for simulation of fracture extension due to the decreased normal stress on
discontinuities, relative to pore pressure. In general, the code has the ability to
simulate the full failure process in small to medium-scale slopes, providing a means to
analyse rock mass and discontinuity strength, along with a representation of the failure
mechanism from initiation through to deposition.

At a large scale the sheer complexity of a fractured rock mass makes it impossible to
model the whole slope as a representative discrete object with an embedded detailed
fracture network. Subsequently an approach is presented in this thesis, whereby one
can use numerical modelling to arrive at a mass strength estimate that can be used in
a simpler equivalent continuum model of a large slope.

Groundwater pressure was initially applied in a simple planar failure model, to provide
confidence in the capability of the newly developed effective stress module within
ELFEN. Following this, groundwater was implemented into two step-path failures.
This highlighted the sensitivity of the specific models to the level of the phreatic
surface, rock-bridge strength and discontinuity related strength. In addition, a fully
drained toe-breakout failure was addressed, using various limit equilibrium and finite

ii
element methods to assess the potential strength of a rock-bridge within the toe of a
50m slope.

In all numerical models it is necessary to be certain of input parameters, or to


understand the implications and effects of any uncertainty. During this thesis an
accumulative scheme or modelling methodology has been followed; starting with
simple models so that comparisons can be made with other limit equilibrium and finite
element methods, allowing calibration of the more advanced properties required within
ELFEN. This calibration is made easier with the use of a staged modelling procedure.
In particular it was found that, when using a dynamic-based code with fracture
capabilities, an inappropriate model procedure can lead to an unrealistic simulation.

In summary, particular contributions and novel aspects of the research were:

i. The application of ELFEN to a variety of scale-related failures in fractured


rock slopes, covering a range of failure mechanisms. In addition, direct
comparisons have been made between the results of ELFEN, limit
equilibrium and finite element methods, for the chosen case study slopes.
This has provided an analysis and initial review of the capabilities and
limitations of each of the individual approaches.
ii. The newly developed groundwater version of ELFEN was tested for the
first time in three of the six case study slopes, demonstrating its
effectiveness in simulating mode I fracture extension and subsequent
slope instability, due to a rise in the phreatic surface.
iii. The development of a suitable staged approach methodology by which a
fractured rock slope can be simulated efficiently and accurately, when
using a dynamic fracture-based code.
iv. The simulation of a large-scale case study slope using a FracMan-ELFEN
approach, whereby a statistically generated fracture network is explicitly
incorporated into a numerical model and mass strength is assessed on a
large scale, deriving strength properties that represent an equivalent
continuum of the fractured mass. Subsequently, a number of approaches
were used to assess the strength of the equivalent continuum that formed
a 1000m slope. This led to the comparison of the numerical and
empirically derived mass strength approach for modelling of slopes.

iii
ELFEN modelling resulted in novel simulations of certain slope failure mechanisms,
which included:

v. The progressive development of a shear plane, resulting in the opening of


a tension crack.
vi. Various step-path failure simulations, demonstrating the sensitivity to both
discontinuity and rock-bridge conditions.
vii. Simulation of an active passive failure mechanism, with subsequent
rotation and translation.

During this thesis, analysis of fractured slopes has been restricted to 2D, due to
computer limitations associated with processing and memory requirements for complex
models. It is anticipated that future developments of a code will be based on parallel
processing, which will allow the use of 3D modelling and provide a platform to consider
larger-scale models of fractured rock slopes.

iv
Contents
ABSTRACT.................................................................................................................... II

CONTENTS.....................................................................................................................v

LIST OF FIGURES ..........................................................................................................xi


LIST OF PLATES ......................................................................................................... xvii
LIST OF TABLES ........................................................................................................ xviii
LIST OF PRINCIPAL SYMBOLS AND ABBREVIATIONS, WITH FUNDAMENTAL UNITS .............. xxi
ACKNOWLEDGEMENTS .............................................................................................. xxiii

CHAPTER 1 – INTRODUCTION.................................................................................... 1

1.1 CONTEXT OF RESEARCH ......................................................................................... 1


1.2 RESEARCH DESCRIPTION ........................................................................................ 3
1.2.1 Aims and Objectives ...................................................................................... 4
1.3 REVIEW OF PAST WORK.......................................................................................... 5
1.3.1 Back-analysis of Randa slope failures using ELFEN ..................................... 5
1.3.2 The application of ELFEN to example slope problems .................................. 5
1.4 MODELLING APPROACH .......................................................................................... 7
1.5 THESIS STRUCTURE ................................................................................................ 9

CHAPTER 2 – LITERATURE REVIEW ....................................................................... 12

2.1 INTRODUCTION ..................................................................................................... 12


2.2 SLOPE STABILITY .................................................................................................. 12
2.2.1 Triggers for rock slope failure....................................................................... 12
2.2.2 Factors that control rock slope failure .......................................................... 13
2.2.3 Effects of Scale ............................................................................................ 20
2.2.4 Modes of fracturing ...................................................................................... 23
2.2.5 Ultimate failure mechanisms ........................................................................ 24
2.2.6 Behaviour of a Progressive Failure .............................................................. 29
2.2.7 Triggers and processes driving progressive-type failure.............................. 31
2.2.8 Examples of progressive failures ................................................................. 33
2.2.9 Numerical modelling of progressive failure .................................................. 34

v
2.3 METHODS OF NUMERICAL MODELLING................................................................... 37
2.3.1 Overview of types of numerical modelling software ..................................... 38
2.3.2 Kinematic Analysis – Siromodel ................................................................... 45
2.3.3 Limit Equilibrium Analysis – SLIDE and other programs used during this
research ....................................................................................................... 46
2.3.4 Hybrid elasto-plastic finite element (boundary) Method – Phase2 ................ 47
2.3.5 Finite Difference (domain) Method – FLAC.................................................. 50
2.3.6 Distinct Element (domain) Method – UDEC and 3DEC ............................... 52
2.3.7 Particle Flow Code ....................................................................................... 55
2.3.8 Hybrid finite/discrete element (domain) Method – ELFEN ........................... 56
2.3.9 Comparison of two-dimensional and three-dimensional analyses ............... 57
2.4 GROUNDWATER WITHIN ROCK SLOPES ................................................................... 59
2.4.1 Influence on overall slope stability................................................................ 59
2.4.2 Groundwater within a fracture-based code .................................................. 64
2.5 METHODS OF ROCK MASS STRENGTH DETERMINATION FOR SLOPES ........................ 65
2.5.1 Classification and chart-based methods ...................................................... 68
2.5.2 Use of numerical modelling for the determination of mass strength ............ 75

CHAPTER 3 – DEVELOPMENT OF A COMPREHENSIVE NUMERICAL MODEL ... 80

3.1 STAGED MODELLING APPROACHES WITHIN FRACTURE-BASED CODES...................... 81


3.1.1 Realistic sequence of loading and release within a slope model ................. 81
3.1.2 Discussion.................................................................................................... 90

CHAPTER 4: CASE STUDY 1 – BENCH-SCALE DISCONTINUITY-CONTROLLED


PLANAR FAILURE ...................................................................................................... 92

4.1 INTRODUCTION ..................................................................................................... 92


4.2 MODEL GEOMETRIES............................................................................................. 92
4.3 CALIBRATION OF MODELLING PARAMETERS ............................................................ 93
4.3.1 Sensitivity of solution to point damping ........................................................ 94
4.3.2 Sensitivity of solution to penalties ................................................................ 97
4.3.3 Sensitivity of solution to further modifications .............................................. 99
4.4 MODIFIED SLOPE, VERSION 4............................................................................... 100
4.5 INTRODUCTION OF A PHREATIC SURFACE AND IMPLICATIONS FOR STABILITY .......... 101
4.5.1 Model without tension crack (planar failure slope 4) .................................. 101
4.5.2 Model with tension crack (planar failure slope 5) ....................................... 104
4.6 SUMMARY AND CONCLUSIONS ............................................................................. 107

vi
CHAPTER 5: CASE STUDIES 2 AND 3 – MODELLING OF MIXED-MASS AND
DISCONTINUITY-CONTROLLED PLANAR FAILURE AT BENCH-SCALE............ 108

5.1 INTRODUCTION ................................................................................................... 108


5.2 CASE STUDY 2 – ‘HUTCHINSON JOSS BAY-TYPE’ CHALK CLIFF FAILURE................. 108
5.2.1 Model Development ................................................................................... 109
5.2.2 Results from limit equilibrium analysis ....................................................... 112
5.2.3 Results from hybrid FEM/BEM (Phase2) .................................................... 115
5.2.4 Results from FEM/DEM (ELFEN)............................................................... 120
5.2.5 Conclusions from Hutchinson Joss Bay Modelling..................................... 128
5.3 CASE STUDY 3 – THEORETICAL STEP-PATH FAILURE THROUGH CHALK .................. 130
5.3.1 Model Geometries ...................................................................................... 130
5.3.2 Results from phreatic surface modelling .................................................... 132
5.4 SUMMARY........................................................................................................... 133

CHAPTER 6: CASE STUDY 4 – MODELLING OF A DOMINANTLY


DISCONTINUITY-CONTROLLED FAILURE AT INTER-BENCH SCALE ................ 135

6.1 INTRODUCTION ................................................................................................... 135


6.2 SIMULATION OF FAILURE MECHANISM ................................................................... 137
6.3 INTRODUCTION OF A PHREATIC SURFACE AND IMPLICATION FOR STABILITY ............ 142
6.3.1 Delabole planar failure modelling ............................................................... 143
6.3.2 Delabole modelling with fracture due to groundwater ................................ 148
6.4 SUMMARY AND CONCLUSIONS ............................................................................. 153

CHAPTER 7: CASE STUDY 5: MODELLING OF A TOE-BREAKOUT MECHANISM


AT KALGOORLIE SUPERPIT................................................................................... 156

7.1 INTRODUCTION ................................................................................................... 156


7.1.1 Key aims of this investigation ..................................................................... 157
7.1.2 Creating a model ........................................................................................ 158
7.1.3 Methods of Analysis ................................................................................... 159
7.1.4 Overview of results..................................................................................... 161
7.2 LIMIT STATE ANALYSES OF ROCK-BRIDGE STRENGTH ........................................... 162
7.2.1 Polygon of Forces ...................................................................................... 163
7.2.2 Multiple slices approach – Jacob ............................................................... 164
7.2.3 Single block – plane failure model.............................................................. 168
7.2.4 Discussion.................................................................................................. 170
vii
7.3 PHASE2 MODELLING ............................................................................................ 171
7.3.1 Phase2 Model 1: Discontinuity Model......................................................... 172
7.3.2 Phase2 Model 2: Non-discontinuity model (breakout through rock mass).. 175
7.4 SUMMARY AND DISCUSSION ................................................................................ 180

CHAPTER 8 – CASE STUDY 6: MODELLING OF A DISCONTINUOUS MASS-


CONTROLLED FAILURE AT LARGE SLOPE SCALE ............................................ 185

8.1 INTRODUCTION ................................................................................................... 185


8.2 FORMATION OF A CASE STUDY SLOPE .................................................................. 186
8.2.1 Locality of previous mapping and basis of model....................................... 188
8.2.2 Previous uniaxial and biaxial ELFEN Modelling ......................................... 189
8.3 ANALYSIS OF MASS STRENGTH ............................................................................ 191
8.3.1 ELFEN Biaxial Tests .................................................................................. 191
8.3.2 Results from the FracMan-ELFEN Approach............................................. 195
8.3.3 Rock Mass strength determined from conventional approaches ............... 205
8.4 MODELLING OF THE FRACMAN-ELFEN EQUIVALENT CONTINUUM ......................... 208
8.4.1 Slide Analysis ............................................................................................. 208
8.4.2 Phase2 Analysis ......................................................................................... 212
8.4.3 Influence of Groundwater on Slope Strength ............................................. 220
8.5 LIMITS FOR SLOPE HEIGHT AND ANGLE FROM DIFFERENT APPROACHES ................. 222
8.5.1 Analysis of downgraded mass strength using FracMan-ELFEN
Approach.................................................................................................... 222
8.5.2 Analysis of mass strength using RocLab-SLIDE Approach ....................... 225
8.6 SUMMARY AND CONCLUSIONS ............................................................................. 229
8.6.1 Selection of DFN ........................................................................................ 229
8.6.2 ELFEN Biaxial Modelling............................................................................ 230
8.6.3 Modelling of the Equivalent Continuum...................................................... 230
8.6.4 Final Summary ........................................................................................... 232

viii
CHAPTER 9: GENERAL DISCUSSION .................................................................... 234

9.1 INTRODUCTION ................................................................................................... 234


9.1.1 Overview of thesis content ......................................................................... 234
9.2 ELFEN CAPABILITIES AND LIMITATIONS................................................................ 238
9.3 BRIEF COMPARISONS BETWEEN ELFEN AND OTHER COMPREHENSIVE NUMERICAL
APPROACHES ...................................................................................................... 239

9.3.1 Phase2 ........................................................................................................ 239


9.3.2 PFC ............................................................................................................ 241

CHAPTER 10: CONCLUSIONS AND RECOMMENDATIONS FOR FURTHER


WORK ....................................................................................................................... 243

10.1 INTRODUCTION ................................................................................................. 243


10.2 CONCLUSIONS .................................................................................................. 243
10.3 RECOMMENDATIONS ......................................................................................... 244

APPENDIX A: IMPORTANT PARAMETERS REQUIRED WITHIN DYNAMIC


FRACTURE-BASED CODE....................................................................................... 247

A.1 MATERIAL PROPERTIES – NON-LINEAR CRITERION .............................................. 248


A.1.1 Fracture Energy ......................................................................................... 249
A.1.2 Fracture Toughness................................................................................... 249
A.2 DISCRETE GLOBAL PROPERTIES AND SURFACE-TYPE PROPERTIES ...................... 250
A.2.1 Damping .................................................................................................... 251
A.2.2 Field ........................................................................................................... 255
A.2.3 Penalties and Stiffness .............................................................................. 256
A.2.4 Zone........................................................................................................... 260
A.2.5 Smallest Element ....................................................................................... 260
A.2.6 Contact Damping Type .............................................................................. 261
A.2.7 Contact Type ............................................................................................. 261
A.3 FACTOR OF CRITICAL STEP ................................................................................. 262
A.4 MESH ASSIGNMENT ............................................................................................ 263
A.4.1 Mesh Density ............................................................................................. 264

ix
A.5 ADDITIONAL EDITS TO THE NEUTRAL FILE ............................................................. 266
A.5.1 Option 166 ................................................................................................. 266
A.5.2 Option 37 ................................................................................................... 267
A.5.3 Option 179 ................................................................................................. 268

APPENDIX:
B - Detail on rock mass criterion used within ELFEN…………………………………..269
C - Slice equilibrium analysis methods within SLIDE…………………………………..272
D - Socio-economic factors that may compromise slope stability…………………….273
E - Empirical-based rock mass characterisation schemes…………………………….274
F - A stress-related issue in ELFEN excavation models……………………………….275
G - Additional results for elfen planar failure models…………………………………...277
H - Suggested modifications for planar failure slope 3 model…………………………283
I - Chalk step-path model 1 (detail on model development and findings)…………….287
J - Data from ELFEN biaxial modelling………………………………………………..….289
K - Derivation of parameters in groundwater models…………………………………...297
L – Model boundaries and mesh details for ELFEN models………………………..….298

MODEL DATABASE……………………………………………………………………….303

LIST OF REFERENCES ............................................................................................ 312

In addition to this thesis, which is presented as one volume, there is one CD-rom
containing a digital copy of the thesis, and the animations of slope failures that are
referred to within the thesis.

x
List of Figures
Page:

Figure 1.1: Case study slopes used during this research………………………………… 8

Figure 2.1: Effects of scale on rock strength (a), demonstrating that a potential failure
mechanism is related to the scale of study (modified after Sjöberg,
1999)…………………………...…………………………………………………. 21

Figure 2.2: Modes of fracture (after van der Pluijm and Marshak, 1997)…..…………... 23

Figure 2.3: Step-path failure with persistent low angle joints (after Stead et al., 2004).. 29

Figure 2.4: Stages of movement that occurs within moving slopes, identified by Fell et
al. (2000 cited in Sullivan, 2007); which can be related to deformation that
occurs during progressive failure…………………………………...…………. 30

Figure 2.5: Decision tree for selecting modelling method (after Scheldt et al., 2003)…. 40

Figure 2.6: Example of fully joint controlled failure mechanisms analysed in UDEC
(after Alejano, 2004)…………………………………………………………….. 54

Figure 2.7: The different areas of mechanics that must be considered when modelling
a slope failure from initiation through to deposition (after Sullivan, 2007)… 61

Figure 2.8: Formation of an equivalent structure, which represents mass strength,


and consequently simulates step-path failure (after Schellman et al.,
2006)……………………………………………………………………………… 67

Figure 2.9: Diagram showing in what situations the Hoek-Brown criterion for mass
strength should be applied (modified after Hoek, 2007)…………………….. 72

Figure 2.10: Illustration of the method by which Mohr-Coulomb shear strength


parameters are derived from a curvilinear Hoek-Brown envelope…………. 73

Figure 3.1: Proposed four-step process by which fractured rock slopes should be
modelled; with component settlement periods highlighted in grey. ……...... 83

Figure 3.2: Dynamic behaviour within an ELFEN model, with displacement of a


discrete block due to successive modelling stages…………….................... 84

Figure 4.1: The three slope models used during the calibration of model parameters... 93

Figure 4.2: Extent of extended boundaries (a) that were used, with other
modifications, in ELFEN planar failure models slopes 4 (a) and 5 (c)…….. 93

Figure 4.3: The influence of point damping (%) upon displacement within ELFEN
planar failure slope 1. (a) Displacement at 1s runtime for a purely
cohesive discontinuity; (b) displacement at 10s for a purely frictional
discontinuity. ……………….. ………………………………………………….. 94

Figure 4.4: The influence of point damping within an ELFEN model of a purely
cohesive discontinuity in planar failure slope 1, with detection of an
appropriate degree of damping………………………………………………… 95

xi
Figure 4.5: Exponential relationship within ELFEN planar failure model 2, which has a
discontinuity strength that dictates a FOS of <1……………………………… 96

Figure 4.6: Influence of mesh density and degree of point damping within ELFEN
planar failure model 2. In addition results from small-scale, relatively
coarsely meshed model 1, is displayed. ………..……………………………. 96

Figure 4.7: Critical degree of penalties (Pn = 0.1Pt) within ELFEN planar failure slope
1, when FOS = 1 (point damping = 10%). (b) Non-uniform behaviour,
discovered when attempting to minimise displacement in a situation where
FOS >1……………………………………………………………………. 98

Figure 4.8: Vertical displacement plot of ELFEN planar failure slope 4 model; note
that where frictional strength is stated, cohesion is zero and vice-versa….. 100

Figure 4.9: Pore pressure distributions within the two fully saturated RocPlane
(Rocscience, 2008) models, which investigate the influence of peak
pressure assignment in a plane failure analysis …...………………………... 102

where discontinuity  = 58° (and c = 0). Average pore pressure is


Figure 4.10: Influence of groundwater within ELFEN planar failure slope 4 model,
2
contoured in N/m and the phreatic surface is also indicated……………… 102

Figure 4.11: (a) Failure when FOS is 1.1 (discontinuity c = 12 kPa,  = 56°), due to
simplified phreatic surface. (b) Prevention of failure by modification of
phreatic surface, in an attempt to achieve similar pressure distribution as
in limit state analysis (c)………………………………………………………… 106

Figure 5.1: Geometry and geology of the Joss Bay model………………………………. 109

Figure 5.2: (a) Failure surface chosen for analysis of the Hutchinson Joss Bay model
in SLIDE (Rocscience, 2008). (b) FOS from non-slice limit equilibrium
method RocPlane, which can only analyse a simplified geometry………… 113

Figure 5.3: Mesh dependency and consequent different behaviour within the
2
Hutchinson Joss Bay Phase model, with a pre-existing tension crack…… 118

Figure 5.4: Model extent (a) and subsequent ability of Phase2 to predict tension crack
development (b) within the Hutchinson Joss Bay model, with no pre-
existing tension crack…………………………………………………………… 119

Figure 5.5: ELFEN Hutchinson Joss Bay model 1, illustrating the meshed regions (a);
development of shear strain when mesh density within the spheres is
0.4m and 0.3m (b) and (c) respectively; (d) addition of ‘Option 37.’……….. 122

Figure 5.6: ELFEN Hutchinson Joss Bay model 2, illustrating the dependency on
tensile strength. (a) (b) Mohr-Coulomb stress paths at the base of the
tension crack, and screenshot of final fractured state, when the σt is
1.5MPa and 0.013MPa respectively. …………………………………………. 123

Figure 5.7: Progressive development of shear strain and subsequent opening of a


tension crack within ELFEN model of Hutchinson Joss Bay cliff section….. 124

Figure 5.8: ELFEN Hutchinson Joss Bay model 3, demonstrating the detection of
failure at point A; which is within the zone of failed elements (a) as shown 124
on Mohr-Coulomb stress path plot (b)…………………………………………

Figure 5.9: Hutchinson Joss Bay unloading models…………………………………….... 125

xii
Figure 5.10: Hutchinson Joss Bay models 4 to 7 (from left to right); (a) runtime, at
which the elements that have reached plasticity intersect the top of the
slope. (b) Final fractured state of each model………………………………. 126

Figure 5.11: (a) (c) Plasticity within ELFEN Hutchinson Joss Bay model 4,
demonstrating the primary detection of compressive failure (b) at point A;
followed by tensile failure at point B as shown in (d)……………………….. 127

Figure 5.12: Analysis of mass strength within an ELFEN Hutchinson Joss Bay model,
demonstrating a shear strength reduction factor of close to 1. Final
degree of plastic strain is plotted for each of the mass strengths…..…….. 128

Figure 5.13: ELFEN models of step-path failure within an example 15m high chalk
cliff: (a) model 1 was restrictive (see Appendix I); (b) no fracture across
rock-bridges occurred in model 2; (c) kinematic release could not be
achieved in model 3; (d) Final model which revealed successful results… 132

Figure 5.14: ELFEN model illustrating instability and consequent step-path failure as
2
a result of rise in pore pressure (displayed in N/m , with the phreatic
surface superimposed). ……………………………………………………….. 132

Figure 5.15: ELFEN model illustrating the final fractured state with low discontinuity
frictional strength and 50% saturation …….…………………………………. 133

Figure 6.1: 1967 failure at Delabole slate quarry, Cornwall; (a) east-west section
from Clover (1978, cited in Coggan and Pine, 1996); (b) conceptual
model illustrating the driving processes……………………………………… 136

Figure 6.2: Original ELFEN model of the Delabole 1967 failure used by Coggan et
al. (2003)………………………………………………………………………… 137

Figure 6.3: (a) (b) (c) Fracturing during respective modelling stages of Delabole
model 1. (d) Shows model 2 where simulation shows kinetic stability,
although fracture does not extend through the rock-bridge………………... 139

Figure 6.4: (a) Geometry of model with extended claylodes (Delabole model 3), to
promote toppling. However fracturing through rock-bridges appears
intensive (b) with no kinematic release………………………………………. 139

Figure 6.5: (a) Step-path geometry, and final state of fracture achieved in model 4
(b). (c) Model 5, which is more kinetically stable, but presents a blocky-
type failure; whereas a slightly different element control and finer mesh in
model 6 (see model database), causes an alternative style of failure (d)... 140

Figure 6.6: (a) Initial step-path geometry when claylodes are extended; (b) results
from model 7. (c) Model 8, where penetration prevents rotation. (d)
Model 9 where different rotations occur when the rock-bridge is finally
removed (e)……………………………………………………………………… 141

Figure 6.7: Simulation of progressive failure through the mass due to a chisel effect,
resulting in rotation and translation, achieved using Delabole model 10… 142

Figure 6.8: Geometry when simulating the Delabole failure as planar mechanism.
Limit state solutions were derived using RocPlane (Rocscience, 2008),
which considers the rear-release surface as a tension crack, with no
particular strength assigned…………………………………………………… 143

Figure 6.9: Fluctuation dynamic-type behaviour indicated through the vertical


displacement after gravity loading, within an ELFEN Delabole planar
failure model where the FOS >1……………………………………………… 144

xiii
Figure 6.10: Over-damped situation, when 10% point damping is used within an
ELFEN Delabole plane failure model, indicated by vertical displacement
of failure block when FOS <1…………………………………………………... 145

Figure 6.11: Kinetic energy associated with releasing a constrained ELFEN Delabole

 = 60°)……………………………………………………………………………
slope model where there is a discrete planar failure block (discontinuity
146

Figure 6.12: (a) Stable situation when a ramped load curve is used within simulation of
Delabole model 12; (b) dynamic-related failure when the same model is
instantaneously loaded…………………………………………………………. 149

Figure 6.13: Saturation of base-plane is critical to stability in ELFEN model 12, where
2
limited pore pressure (N/m ) upon base-plane allows stability (a). Failure
occurs in the same model, when whole length of base-plane is saturated
(b)…………………………………………………………………………………. 150

Figure 6.14: Progressive failure of rock-bridge within ELFEN step-path Delabole model
12, as a result of a rise in pore pressure…..………………………………….. 151

Figure 6.15: Progressive failure of rock-bridge, as a result of a rise in pore pressure,


within ELFEN step-path Delabole model 13. Full saturation of base-plane
and shortah (to 50% height), occurs at 10s runtime, resulting in fracture of
rock-bridge, although further review of block interaction is required………. 152

Figure 7.1: East-west cross section through Brownhill (taken at 49350m north);
illustrating the failed areas, lithologies and major discontinuities…………... 158

Figure 7.2: Annotated cross section through Brownhill, illustrating the actual and
simplified profile, which can be split into 21 slices for analysis in Jacob….. 162

Figure 7.3: Polygon of forces approach to analysing the cohesive strength of the
rock-bridge required to prevent the mass from sliding………………………. 163

Figure 7.4: Profile of Brownhill Jacob Models 1 and 2, with annotation illustrating the
slices from which the high horizontal force originates in Jacob Model 1….. 163

Figure 7.5: Plane failure model for Brownhill, where the cohesive strength of the rock-
bridge can be derived used two different approaches………………………. 167
2
Figure 7.6: (a) Primary Phase Model 1, with degree of straining presented; (b) same
model but extended boundaries causing a reduction in straining in the
region close to the Subvertical Fault…………………………………………... 172

Figure 7.7: Comparison between results from same Phase2 Model 2, but with
different mesh densities; strain is contoured on same interval for both (a)
and (b)…………………………………………………………………………….. 175

Figure 7.8: Development of shear strain (in case 6 within Table 7.11) along rock-
bridge and consequent yielding upon discontinuities; due to excavation
and strength reduction within the rock-bridge………………………………… 178

Figure 7.9: Summary of results, from the different techniques used in the back-
analysis of the potential cohesive strength for the rock-bridge at toe of
Brownhill………………………………………………………………………….. 181

xiv
Figure 8.1: Sequence within the FracMan-ELFEN (numerical) approach, used to
derive and analyse rock mass strength in a large slope; an empirical-
based method is also presented for comparison…………………………….. 186

Figure 8.2: Illustration of how large-scale ELFEN biaxial models (incorporating a


DFN) are used to derive slope mass strength represented by an
equivalent continuum. Zoning of mass strength is included, to allow the
analysis of changing shear strength with depth……………………………… 188

Figure 8.3: DFN for a pillar in Middleton Limestone Mine, with presentation of the
derived jointing on pillar faces (after Elmo et al., 2005)……………………... 189

Figure 8.4: Typical mesh used in the ELFEN biaxial models, with annotation marking
the axis along which stress monitoring points were located …..…………… 194

Figure 8.5: Large (14 x 7m) Middleton mine pillar model from Elmo (2006), from
which a horizontal section is extracted and rotated to create an ELFEN
biaxial model with a dominantly horizontal fracture pattern………………… 194

Figure 8.6: Fractured state of horizontal section when model is run with default
fractures (a); fractures highlighted in (b) are removed to allow the
predominant fracture orientation to control failure…………………………… 195

Limestone, derived from peaks in 1 which occur at low strain. (b)


Figure 8.7: (a) Presents data that marks the onset of failure in the Hoptonwood

Presents the yield behaviour of the respective sections, which occur at


higher degrees of axial strain…………………………………………………... 196

material, derived from peaks in 1 which occur at low strain. (b) Presents
Figure 8.8: (a) Presents data that marks the onset of failure within downgraded

the yield behaviour of the respective sections, which occur at higher


degrees of axial strain………………………………………………….............. 199

Figure 8.9: Stress-strain plot of downgraded RE biaxial models, with data from a
4MPa confined simulation; screen shots illustrate the state of fracture at
which the strengths for the respective models were taken…………………. 200

Figure 8.10: (a) Data from downgraded RE model, demonstrating an exponential


envelope upon a Hoek-Brown plot; consequently a Mohr-Coulomb failure
envelope is more appropriate as presented in (b)…………………………… 201

Figure 8.11: Synthesis of all data derived from ELFEN biaxial testing. (a) and (b)
Present the failure envelopes from low axial strain; (c) and (d) indicate
the results when the mass strength is taken from a higher axial strain…… 204

Figure 8.12: Data from ELFEN testing on non-downgraded RC and RE biaxial sections
forming average Hoek-Brown envelopes and Mohr-Coulomb envelopes
respectively in (a) and (b) respectively………………………………………... 205

Figure 8.13: Illustration of the composition of a ‘zoned’ SLIDE model, to allow for the
rotation of biaxial stress with depth (relative to the shear strength on a
large circular failure surface)…………………………………………………… 209

Figure 8.14: Shallow circular failure within a SLIDE model of a 45° slope with Mohr-
Coulomb mass properties c = 164kPa and  = 58°…………………………. 210

xv
2
Figure 8.15: Mesh and extent of model used in the Phase 55° large slope simulations.
Annotation is given in the zoomed view to indicate the extent of failure
when SRF = 1.31…………………………………………………….................. 214

Figure 8.16: Mohr-Coulomb plots of stresses at two points at different stages during
the excavation and subsequent SSR of a 1000m slope composed of a
mass strength from the downgraded (vertically fractured) Middleton
DFN……………………………………………………………………………….. 215

Figure 8.17: (a) Mohr-Coulomb plot of stress at Point B during the excavation of a
1000m slope composed of a mass strength derived from a SRF of 1.31 on
the downgraded (vertically fractured) Middleton DFN. (b) Stress path
derived from the Mohr-Coulomb plot………………………………………….. 216

Figure 8.18: (a) Mohr-Coulomb plot of stress at Point C during the excavation of a
1000m slope composed of a mass strength derived from a SRF of 1.31 on
the downgraded (vertically fractured) Middleton DFN. (b) Stress path
derived from the Mohr-Coulomb plot………………………………………….. 217

Figure 8.19: (a) Shear strain and failed nodes, during the final three stages of
excavation of a 1000m slope. (b) Maximum principal stress and failed
nodes of the same model and stages as in (a), with scaled stress tensors
to illustrate the orientation and ratio between σ1 and σ3…………………….. 219

Figure 8.20: Comparison of critical failure planes predicted by SLIDE and the Hoek
and Bray (1981) circular failure charts, when different groundwater
conditions are assumed………………………………………………………… 221

Figure 8.21: Comparison of data from RC and RE biaxial tests with the results that
Elmo (2006) found from testing RA, which has the same fracture intensity. 226

Figure 8.22: DFN within models RA, RC and RE, which all have the same fracture
intensity (P21) and GSI. However fracture orientation differs causing a
significantly weaker mass strength in RC…………………………………...... 226

Figure 8.23: Data from downgraded FracMan-ELFEN tests. The horizontally aligned
Middleton DFN, presents a high mass strength that cannot be considered
using mi suggested in RocLab…………………………………………………. 228

xvi
Page:
List of Plates
Plate 7.1: (a) Panoramic photograph (looking south) of the Fimiston open pit,
Kalgoorlie, highlighting the failure area; (b) enlarged view of the failed zone.
(c) View from the west (looking north east), with delineation of failed
zone…………………………………………………………………………………. 157

Plate 8.1: Typical view of Middleton Limestone Mine in Derbyshire (after


Elmo, 2006)………………………………………………………………………….
188

xvii
List of Tables Page:

Table 1.1: Previous applications of ELFEN within literature, applied to theoretical


slope problems…………………………………………………………………... 6

Table 2.1: Types of damage that can occur within slope masses, with the potential to
ultimately lead to slope failure (Stead et al., 2004)………………………….. 13

Table 2.2: Basis composition of some of the more frequently used methods of slope
stability analyses that are available, with suitability, benefits and
limitations…………………………………………………………………………. 41

Table 2.3: Derivation of the Hoek-Brown constants for disturbed and undisturbed
rock masses, based on RMR. (Equations from Hoek and Marinos, 2006).. 70

Table 3.1: Methods of slope restraint within ELFEN, with specific problems
suggestions and disadvantages to each technique………………………….. 89

Table 4.1: Results from simulations on planar failure slope 3, demonstrating the
sensitivity of the solution to variation in penalties assigned to the single
discontinuity defining the discrete failure block …..………………………….. 99

Table 4.2: Analysis of planar failure slope 4, proving ELFEN to conform to simple
limit states predicted by RocPlane…………………………………………….. 103

Table 4.3: Analysis of planar failure slope 5, with failure occurring in ELFEN when
Plane_failure predicts a FOS >1, due to a simplified phreatic surface……. 105

Table 4.4: Analysis of planar failure slope 5, where the shape of the phreatic surface
has been modified in the ELFEN models, as indicated by 4.11b…………... 106

Table 5.1: Peak strengths for the Chalk at Joss Bay, reported by both Hutchinson
(1970 cited in Hoek and Bray, 1981) and Hoek and Bray (1981)………….. 110

Table 5.2: Strength inputs for limit equilibrium analysis of mass strength in
Hutchinson chalk cliff model……………………………………………………. 113

Table 5.3: Summary of results from different limit equilibrium analyses, using the
properties and geometries presented in Table 5.2 and Figure 5.2………… 114

Table 5.4: Derivation of mass tensile strength for chalk using RocLab (Rocscience,
2008) and the Mohr-Coulomb mass strength, suggested by Hoek and
Bray 1983)………………………………………………………………………... 116
2
Table 5.5: List of material properties, used within the Phase simulation of the
Hutchinson Joss Bay model……………………………………………………. 117

Table 5.6: Basic material and discontinuity strength properties used in chalk step-
path models………………………………………………………………………. 130

Table 6.1: Summary of the most important properties used for the mass and
discontinuity strength in Delabole models…………………………………….. 138

Table 6.2: Results from a number of Delabole wet-slope plane failure simulations,
where different limiting conditions were applied to the discontinuities
defining discrete failure block…………………………………………………... 147

Table 6.3: Summary of the most important properties used for the mass and
discontinuity strength in Delabole model 12………………………………….. 148

xviii
Table 6.4: Results from simulations on Delabole step-path model 12, where the
phreatic surface was raised to different heights, and/or discontinuity
frictional strength was altered to cause failure……………………………….. 152

Table 6.5: The occurrence of premature rock-bridge failure within ELFEN Delabole
step-path model, with prevention methods and recommendations………… 155

Table 7.1: Basic construction of the five different limit equilibrium models used for
the analysis of cohesive strength of the rock-bridge; with advantages and
disadvantages listed for each technique……………………………………… 160

Table 7.2: Basic construction of the two different finite element models used for the
analysis of cohesive strength of the rock-bridge; with advantages and
disadvantages listed for each technique……………………………………… 161

Table 7.3: Results of a sensitivity study on the strength of the rock mass that forms
the rock-bridge in Jacob Model 1……………………………………………… 165

Table 7.4: Results of a sensitivity study on the strength of the rock mass that forms
the rock-bridge in Jacob Model 2…...……………………………………….… 167

Table 7.5: Results from sensitivity study using Plane_Failure.xls (Pine, 2006b), with
resulting cohesive properties of the rock bridge, when calculated from the
tension required in a rock bolt………………………………………………….. 169

Table 7.6: Results from sensitivity study using Plane_Failure.xls (Pine, 2006b), with
resulting cohesive properties of rock-bridge calculated from the cohesive
strength required on the base plane to prevent the mass from failing…….. 169

Table 7.7: Summary of results from the sensitivity study of rock-bridge cohesive
strength, using different limit equilibrium models…………………………..... 170
2
Table 7.8: Phase (Model 1) study of rock-bridge strength. Note that the analysis
was performed sequentially, with the parameter which has been altered
between each case highlighted in bold……………………………………….. 174

Table 7.9: Influence of mesh density within (un-staged) Phase2 Model 2 of the
Brownhill failure………………………………………………………………...... 175

Table 7.10: Results of a sensitivity study on the strength of the rock mass that forms
the rock-bridge in Phase2 Model 2 (using a default k ratio of 1)……………. 177

Table 7.11: Results of a sensitivity study on the strength of the rock mass that forms
the rock-bridge in a Phase2 Model 2 with a k ratio of 2.32………………….. 179

Table 7.12: Situations used to compare the different methods of analysis that have
been used during the investigation of rock-bridge strength in the toe of
Brownhill. Methods from which results can be drawn are presented……… 180

Table 7.13: Comparison of results from selected methods of analysis, which can
consider frictional strength, on the rock-bridge within the toe of Brownhill... 182

Table 8.1: Summary, collated by Z. Flynn (2007, personal communication), of results


from the simulations of Middleton Mine pillars presented in Elmo (2006)… 190

Table 8.2: Properties for ELFEN modelling parameters for biaxial simulations,
derived from the uniaxial simulations carried out by Elmo (2006)…………. 192

Table 8.3: Hoek-Brown and Mohr-Coulomb properties from non-downgraded RC,


RE and horizontal ELFEN biaxial models…………………………………….. 197

xix
Table 8.4: Original (from Elmo, 2006), and quartered, intact strength properties for
ELFEN biaxial models with the later being for the weaker/downgraded
models…………………………………………………………………………… 198

Table 8.5: Hoek-Brown and Mohr-Coulomb properties from RE and horizontal


ELFEN biaxial models on the downgraded Hoptonwood Limestone……. 199

Table 8.6: Mohr-Coulomb properties from downgraded RE and Horizontal FracMan-


ELFEN biaxial models………………………………………………………..... 202

Table 8.7: Calculation and consequent FOS from the Hoek and Bray circular failure
charts (1981) on a dry 55° 1000m dry slope; note that in this case F
refers to factor of safety, and H to slope height……………………………... 202

Table 8.8: Strength inputs for Hoptonwood Limestone in Middleton Mine, obtained
from Pine et al. (2006)…………………………………………………………. 206

Table 8.9: Range of Hoek-Brown mass strength parameters for the Hoptonwood
Limestone, which have been derived using RocLab (Rocscience, 2008),
for comparison against results from the FracMan-ELFEN analysis, listed
in Table 8.10............................................................................................... 206

Table 8.10: Range of Hoek-Brown mass strength parameters for the Hoptonwood
Limestone, that have been derived using the FracMan-ELFEN
approach……………………………………………………………………….... 206

Table 8.11: Results from SLIDE modelling of Mohr-Coulomb mass strength derived
from the ELFEN biaxial testing………………………………………………... 211

Table 8.12: Results from SLIDE modelling of Hoek-Brown mass strength derived
from the ELFEN biaxial testing………………………………………………... 211

Table 8.13: Results from a Shear Strength Reduction (SSR) study on a 55° 1000m
high slope; with resulting Shear Strength Reduction Factor and
consequent mass properties which can be checked in a SLIDE model….. 213

Table 8.14: Calculus and consequent FOS from the Hoek and Bray circular failure
charts (1981); note that in this case F refers to factor of safety…………… 220

Table 8.15: Influence of groundwater on the strength of a 55° 1000m slope


composed of material derived from ELFEN biaxial modelling on vertically
aligned Middleton DFN)……………………………………………………...... 220

Table 8.16: Comparisons between different criteria used to derive and analyse the
mass strength from the FracMan-ELFEN approach………………………... 224

Table 8.17: Calculus and consequent FOS from the Hoek and Bray circular failure
charts (1981) for a 1000m slope……………………………………………... 225

Table 8.18: Rock mass strength for the downgraded material, derived directly from
intact strength using the empirical Generalised Hoek-Brown Criterion…... 226

Table 8.19: Rock mass strength for the downgraded material, derived directly from
intact strength using the empirical Generalized Hoek-Brown Criterion
with a higher GSI..………............................................................................ 229

Table 8.20: Summary of methods, advantages, disadvantages and comparisons


between each approach used in the analysis of the large slope case
study……………………………………………………………………………… 231

xx
List of principal symbols and abbreviations, with fundamental
units (where applicable)

a: Hoek-Brown criterion constant


avi : Audio video interleave file
c : Cohesive strength (Pa)
D: Hoek-Brown criterion constant (disturbance factor)
DFN: Discrete fracture network
1 : Maximum displacement (m)
E : Young’s modulus (Pa)
FOS: Factor of safety
G : Energy release rate (J/m2)
Gf : Fracture energy (J/m2)
GSI: Geological Strength Index
Kn : Joint normal stiffness (Pa/m)
Ks : Joint shear stiffness (Pa/m)

KIC : Fracture Toughness (Pa m )


m : Hoek-Brown criterion constant
mi : Hoek-Brown criterion constant
mb : Hoek-Brown criterion constant
μ : Pore pressure (Pa)
P20 : FracMan notation for number of fractures per unit area of rock mass (m-2)
P21 : FracMan notation for length of fractures per unit area of rock mass (m/m2)
P32 : FracMan notation for area of fractures per unit volume of rock mass (m2/m3)
Pn : ELFEN normal penalty coefficient (Pa/m)
Pt : ELFEN tangential penalty coefficient (Pa/m)
 : Frictional strength (degrees)
R : Rayleigh damping
R2 : Coefficient of determination
RMR: Rock mass rating
s: Hoek-Brown criterion constant
SRM: Synthetic rock mass
SSR: Shear strength reduction
SRF: Stress reduction factor
1 : Major principal stress (Pa)
3 : Minor principal stress (Pa)
t : Tensile strength (Pa)

xxi
c : Uniaxial compressive strength of the rock mass (Pa)
ci : Uniaxial compressive strength of intact rock (Pa)
H : Maximum horizontal principal stress (Pa)
h : Minimum horizontal principal stress (Pa)
σn : Normal stress (Pa)
σn’ : Effective normal stress (Pa)
v : Vertical principal stress (Pa)

 : Shear strength (Pa)


Poisson’s Ratio
pc : Failure plane inclination (degrees)
f : Dip of slope face (degrees)
 : Unit weight (N/m2)

xxii
Acknowledgements
I am particularly thankful to both my supervisors Prof. Robert Pine and Prof. John
Coggan for their help and collaboration during this research.

Thanks must go to all of my family and friends who provided words of encouragement;
especially to my partner Zoe Jarvis for her support even throughout her studies, and
also to my parents who have provided a home perfect for focussed research.

Training and code developments for ELFEN must be credited to Rockfield Software
Ltd. They also provided support during the progress of the research project; for this
I am specifically appreciative of the efforts of Melanie Armstrong, Jon Rance, Mauricio
Lobao and Prof. Roger Owen.

Collaboration throughout this project has also been from other academic and industrial
partners, including the Kalgoorlie Consolidated Gold Mines, University of Swansea,
Golder Associates, IMERYS, Rio Tinto and University of Brighton.

The research outlined in this thesis is part of an Engineering and Physical Sciences
Research Council (EPSRC, Grant No. EP/C518713/1) project at Camborne School of
Mines (CSM), University of Exeter. Part-funding was also gratefully provided by the
Isle of Man Government, who have provided financial support with the tuition fees, not
only in the doctoral research outlined in this thesis, but throughout previous bachelor
studies.

I would also like to thank the Camborne School of Mines Trust and the Institute of
Materials, Minerals and Mining for the award of travel scholarships that allowed me to
travel and attend two international conferences. In addition these funds also provided
the opportunity of visiting the Fimiston Open Pit in Western Australia, which formed a
case study due to the supportive geotechnical staff at KGCM.

Finally during the PhD, research has been undertaken in the presence of associated
colleagues whom collectively formed the Geomechanics Group, at Camborne School
of Mines. Thanks must go to them for inspiration and collaboration.

xxiii
CHAPTER 1: Introduction

Chapter 1 – Introduction

Chapter 1 introduces the context of the research, outlining research aims and
objectives. There is a summary of the past work that has been completed on fractured
rock slopes, using a comprehensive fracture-based code ELFEN, which is the main
geomechanical software that has been used throughout this thesis. The review of past
ELFEN work is split into two areas: the application of ELFEN to the back-analysis of
large slope failures, and the application of ELFEN to example slopes. An approach to
the research is presented through the consideration of six case study rock slopes,
encompassing three increasing scales of failure. Finally there is an introduction into
the thesis structure, outlining the content of each of the Chapters within this thesis.

1.1 Context of Research


In order to deliver to a growing demand for commodities, modern mining is tending
towards large-scale operations. The recent global trend is to exploit relatively low
grade competent ore bodies, developing large open pits with minimal footprint; also it
appears the current movement is to implement more block-cave operations, with some
excavating simultaneously alongside open pit mining. To allow efficient extraction
under these conditions, the understanding of rock strength is imperative. However, the
prediction of rock behaviour can be an in-depth and difficult task due to the
unpredictable nature of rock, with many variables all of which influence its strength at
different scales. Like many others, Board et al. (1996) state: “Rock is a highly complex
material that often behaves in a non-linear and anisotropic manner.”

Traditionally, mine-design decisions concerning rock-mechanics have been made with


reference to experience, however with the increasing scale of modern mines, rock
mass behaviour is moving into areas beyond the empirical knowledge-basis; therefore
the understanding of rock strength needs to increase and improve. Mas Ivars et al.
(2007) suggest that joint fabric, post-peak behaviour and scale effects are areas within
geomechanics that are poorly understood. In addition, with the deepening of open pits,
purely structurally controlled failures become less frequent; the focus is instead shifted
to more complex failure mechanisms that may have the ability to compromise large-
scale wall integrity. The understanding of such mechanisms is relatively limited and
requires coupling of many processes; for instance firstly the DIANE (discontinuous,
inhomogeneity, anisotropy and non-elastic) features of a rock mass need to be
considered.

1
CHAPTER 1: Introduction

Numerical modelling is widely used as a tool during the analysis of rock mass strength
at various scales; however there are many different methods available. This research
therefore reviews mechanisms of rock slope instability and the current methods of
numerical modelling that may be used. Subsequently the main objective is to build on
the understanding of complex rock strength, through modelling example and actual
rock slopes in a range of software, with emphasis on furthering the application of one
particular comprehensive approach, ELFEN (Rockfield, 2008).

It is important to respect that one can only ever regard findings from models as an
indication of potential mechanisms. In addition, the data on which to base a
comprehensive model is often not routinely collected. Therefore empirical knowledge
should be considered alongside numerical models, with a view towards to the
development of Internet-based slope databases where case studies of larger slopes
from around the World could be recorded as an effort to update the now restrictive
empirical approaches.

Techniques available for monitoring slope behaviour have improved significantly over
the last decade with advanced tools now widely accessible, but costly. In some
respects this discourages the development of advanced numerical modelling, as
mining companies often have relatively short term interests and can rely more on
monitoring, with appropriate design development as problems arise. In a review of
slope design for the east wall at Mantovrede Mine in Chile, Schellman et al. (2006)
suggest that slope design can now be “flexible” with basis on observational methods.
In particular they refer to results from limit equilibrium and other continuum based
numerical analysis, being used to create a contingency plan to follow if a sudden
increase in displacement rate is observed during the development of steeper slopes on
the east wall of the pit.

Instead of discouraging the development of advanced numerical modelling, mining


companies could benefit, with the potential of evaluating model performance from the
advanced methods of slope monitoring that are available. Currently there appears to
be limited research on the integration of slope monitoring and the comprehensive forms
of numerical modelling; possibly due to the complexity of the numerical codes and their
limited application to fractured rock slopes. Consequently this thesis introduces
various applications of a dynamic fracture-based numerical modelling software,
allowing the development of an approach that should be followed when simulating a
fractured rock slope.

2
CHAPTER 1: Introduction

Finally, the ability of numerical modelling to evaluate slope failure mechanisms, and
rock mass strength, is explored during the research outlined in this thesis. This has
direct applications in the excavation of open pit mines, improving efficiency and safety.
Assessment of rock slope stability is also especially important in a range of disciplines
beyond that of mining. It can be considered that the development of numerical
modelling is imperative to furthering the understanding of fundamental rock mechanics,
although it is of utmost importance to ensure numerical modelling is never the sole
basis of an investigation, instead it should remain only as a tool within a larger scheme.

1.2 Research description


There are limitations involving numerical modelling software, and the simulation of the
full processes of failure within fractured rock slopes. Conventional continuum-based
methods are only able to model the elastic/plastic deformation, whereas discontinuum
approaches are only able to simulate failure on pre-existing discontinuities without the
appreciation of the fracture extension and the initiation of new fractures. The use of the
hybrid finite-element method (FEM)/discrete-element method (DEM) software, ELFEN,
allows the modelling of brittle fracture initiation, propagation and the comminution of the
discrete elements that are formed. This permits the total slope-failure process to be
modelled from initiation with the progressive development of a large-scale failure
surface, upon which kinematic release can be simulated, resulting in the transport and
deposition of the failure block (Eberhardt et al., 2003; Eberhardt et al., 2004; Stead and
Coggan, 2006). The primary aim of this research is to assess the application of
ELFEN, to slope failure processes in fractured rock. In addition various limit
equilibrium-based solutions, and a finite element code (Phase2), were also used during
analysis of the case study slopes; subsequently different approaches are evaluated.
Furthermore a key part of the research is to evaluate the newly developed groundwater
module within ELFEN, via simulating slope instability situations within fractured rocks
that are under effective stress.

During the PhD study, various field work activities were undertaken with an aim of
integrating remote data capture with numerical modelling; the intention was to use this
research aspect to optimise the sequence of data capture, fracture network modelling
and numerical modelling. However this aspect of research could not be concluded due
to time constraints and certain areas of the collaborative work being incomplete.

3
CHAPTER 1: Introduction

1.2.1 Aims and Objectives


The following aims and objectives summarise the research efforts presented within this
thesis:

Aims
1. Evaluate the ability of a dynamic fracture-based code (ELFEN), to simulate rock
slope failures through blocky rock masses.
2. Model slope instability examples within fractured rocks that are under effective
stress and thereby assess the newly developed groundwater module within
ELFEN, with comparisons against other finite element and limit equilibrium
programs.
3. Evaluate the combined use of FracMan and ELFEN as a means to determine
rock mass strength.

Objectives
1. Review present techniques to determine rock slope mass strength.
2. Develop a suitable approach to establish and unload a fractured slope model
within a dynamic fracture-based code (ELFEN).
3. Use different modelling methods to validate behaviour predicted by ELFEN.
4. Contribute to the understanding of complex rock strength through the analysis
of example and actual rock slope failures at a variety of scales. Importantly a
range of failure mechanisms should also be considered using different
modelling approaches, from simple small-scale discontinuity-controlled bench
scale, to larger overall rock mass-controlled slope failures.
5. Calibrate numerical models against known situations (intact geomechanical
tests and back-analysis of slope failures).
6. Use numerical models to determine rock mass strength.
7. Further the application of numerical modelling to rock mass strength
determination, via the use of the ‘FracMan-ELFEN’ approach (Pine et al., 2006),
as introduced in Section 2.5.2.
8. Subsequently present recommendations for potential improvements to the
traditional empirical methods of rock slope strength and stability appraisal.

4
CHAPTER 1: Introduction

1.3 Review of Past Work


Recently ELFEN was used by Elmo (2006), to investigate its application to
underground hard rock pillars, thereby testing the ability of ELFEN to simulate
displacement and rotation along existing natural joints, and failure within intact rock
material. Coggan et al. (2003) documented how ELFEN has been applied successfully
to the simulation of breakouts and rock bursts within deep mines; blasting of rock
slopes in open pit mining; simulation of rock tests; the mining of tabular orebodies and
even comminution and grinding during mineral processing.

ELFEN has been applied to many dry rock slope problems, highlighting its ability to
capture the complete slope failure from initiation to transportation and deposition.
Section 1.3.1, presents situations where ELFEN has been used to back-analyse large
slope failures. Section 1.3.2 presents some of the example applications from literature
to which ELFEN has also been applied.

1.3.1 Back-analysis of Randa slope failures using ELFEN


The implementation of ELFEN as a tool to back-analyse the 1991 Randa failure in
Switzerland, demonstrated how the software could simulate progressive failure and the
total slope failure process (Eberhardt, 2003; Stead and Coggan, 2006). The modelling
of the Randa failure was also documented within Eberhardt et al. (2004); who
suggested the importance of cumulative damage and fracture due to cyclic fatigue,
which may have allowed strength degradation. This unloading during deglaciation has
been shown to have resulted in tensile damage at the slope toe (Stead and Coggan,
2006). The consequence of precipitation and the snow-melt, prior to failure, provided a
final impetus to allow the strength degradation to reach its equilibrium state and
subsequent failure to occur. Also Eberhardt et al. (2004) used the progressive nature
along with mechanisms involved in the Randa failure, to explain the reports of tilting
and falling of large blocks that occurred during the failure, which progressed over a
number of hours. They also suggested progressive failure mechanisms to be the
reasoning behind the relatively short runout of debris.

1.3.2 The application of ELFEN to example slope problems


To date, numerous applications of ELFEN have taken place. Table 1.1 summarises
these applications and some of the findings.

5
CHAPTER 1: Introduction

Table 1.1: Previous applications of ELFEN within literature, applied to theoretical slope problems.

Failure Mechanism Reference Detail of Model Findings

Fragmentation of the weak rock mass


Planar failure Stead et al. (2004) (tensile strength = 1 MPa), as the block
After Stead et al. (2004)
slides down slope.

Coggan et al. (2003) and Completion of a low angle joint via the
Step-path failure
Stead et al. (2004) After Stead After Coggan propagation of cross over fractures
et al. (2004) et al. (2003)

Coggan et al. (2003); Translational sliding due to gradual


Active-passive wedge
Stead et al. (2004); and development of fractures within the rock
(bi-planar failure) After Stead
Stead & Coggan (2006) mass
et al. (2004)

Occurring in thinly
Footwall (thin-bedded) bi-
Stead et al. (2006) and bedded tabular bodies Simulation of tensile fracturing,
planar/ Buckling and
Stead et al. (2004) After Stead and low angle shears providing kinematic release
ploughing type failure
et al. (2004)

Coastal instability
Coggan et al. (2003) and due to undercutting Vertical breakdown of material due to
Buckling and toppling
Stead & Coggan (2006) the initial undercut
After Coggan
et al. (2003)

Simulation of full process, with


After Stead et
interlayer slip and flexure of the rock
Flexural toppling Stead et al. (2004)
al. (2004) columns as they bend forward and
consequent fracture
6

6
CHAPTER 1: Introduction

During the research period, associated research within CSM was concerned with the
simulation of the fracturing process of pillars and during block caving operations. This
implemented the FracMan-ELFEN interface, with FracMan modelling at CSM (Ford,
2008), and the subsequent ELFEN modelling at Rockfield. Linking the FracMan
modelling to methods of remote data-capture is another related area of research within
CSM, with the investigation of Sirovision (CSIRO, 2008a) to capture discontinuity data
from digital photography of both surface and underground rock faces (Gwynn, 2008).

1.4 Modelling Approach


To test the applicability of both current and new effective stress versions of ELFEN, six
case study slopes have been modelled to include a variety of failure mechanisms (both
mass and discontinuity controlled). In addition, these example slopes encompass
three different scales of study, small (< 20m), medium (20 to 100m) and large-scale
(>100m).

Case study and example slopes:


1. (Figures 1.1a and 1.1b) Small-scale example discontinuity-controlled failures,
simulating sliding upon a planar failure both without and with a tension crack.
2. (Figure 1.1c) Small-scale case study model of a chalk cliff in NE Kent, where
dominantly mass-controlled failure occurred via the progressive development of
a shear plane.
3. (Figure 1.1d) Small-scale example model of chalk cliff where mixed
discontinuity and mass-controlled failure occurs following a step-path
mechanism.
4. (Figure 1.1e) Medium-scale dominantly discontinuity-controlled case study
failure at Delabole Slate Quarry, Cornwall, simulating development of an active-
passive wedge mechanism and investigating progressive failure across rock-
bridges.
5. (Figure 1.1f) Medium-scale mixed discontinuity and mass-controlled case study
failure at Fimiston Open Pit (Kalgoorlie Gold Mine), Australia. The principal aim
of modelling was to back-analyse cohesive strength of mass where toe
breakout occurred.
6. (Figures 1.1g and 1.1h) Study of example fractured rock slope behaviour at
large-scale. Biaxial data, based on blocky rock mass mapping, was used to
derive rock mass strength. This was then simulated as an equivalent
continuum, through which large-scale mass-controlled failure occurs.

7
CHAPTER 1: Introduction

Case studies 1 to 4 were used to test the new groundwater module within ELFEN;
therefore simulations were completed with and without the pressures induced by
groundwater within the discontinuities. As a result of the nature of the data that was
available from Kalgoorlie, case study 5 was not analysed in ELFEN. Also due to time
constraints, ELFEN was only used for the primary (biaxial) analysis within the final
large-scale slope in case study slope 6.

Figure 1.1: Case study slopes used during this research.

8
CHAPTER 1: Introduction

The six case studies presented in Figure 1.1, are used to evaluate the suitability of
different numerical modelling methods, to the simulation of rock slope failures from the
simple bench-scale (case studies 1 to 3) through to the more complex staged
excavation of intermediate-scale slopes (case studies 4 and 5). The final scale of
study (case study 6) investigates the derivation of an equivalent continuum from
comprehensive biaxial modelling, which is subsequently modelled as a large-scale
slope using a variety of more simple analytical approaches, and numerical methods.

During the analysis of most of the case studies, different modelling methods are used
alongside ELFEN, allowing comparison and validation of a variety of numerical
approaches. In particular the analysis method in case study 5, uses less
comprehensive numerical modelling; instead investigation progresses from limit
equilibrium approach to the more comprehensive finite element model, to determine
potential rock mass strength within the toe region.

1.5 Thesis Structure


This thesis is presented through ten Chapters. The present Chapter provides an
introduction to the research context, aims and subsequent approach; and the final
Chapter concludes the research and main findings with recommendations for further
work.

The Appendix to this thesis has numerous sections, most of which detail the specifics
of different aspects of the numerical modelling software used, and in some cases the
raw results from these. Following the Appendices, there is a Microsoft Excel
spreadsheet (referred to as the ‘model database’) summarising the detail for each of
the numerical models referred to throughout this thesis.

In addition a CD-rom is enclosed, which contains the following:

 Digital copy of this thesis


 The model database
 Audio Video Interleave (avi) files of specific simulations referred to within the
text.

As discussed in the previous section, the structure of thesis is mainly based around six
case studies, each of which investigates an aspect of slope strength, but on an
increasing scale. A summary of the content of Chapters 2 to 8 is given below.

9
CHAPTER 1: Introduction

Chapter 2 – Provides an in-depth literature review, split into five parts each based upon
a specific area that had to be considered during this thesis. Primarily there
is an introduction, followed by a section which reviews the factors that
influence slope stability, and presents the failure mechanisms that can occur
within fractured rock slopes. The third section outlines methods of
numerical modelling, from simple to complex, which can be applied to
fractured rock slopes. The basis of each approach is detailed with a
comprehensive review of the subsequent advantages and disadvantages of
each numerical method. The influence of groundwater is discussed within
the fourth section, reviewing aspects that can be considered using the newly
developed groundwater module within ELFEN. Finally the fifth section
introduces methods of assessing mass strength, outlining new numerical-
based techniques that can be compared with conventional empirical based
approaches, as applied during Chapter 8.
Chapter 3 – Introduces aspects that need to be considered when developing a model
of a fractured rock slope, within a dynamic fracture-based numerical code.
A procedure is outlined, which should be followed when using a
comprehensive dynamic numerical model to simulate failure within a
fractured rock slope. Consequently different approaches within each of the
modelling stages during this procedure are discussed, leading to
recommendations and detail on appropriate techniques during the modelling
procedure.
Chapter 4 – Further familiarisation of the numerical parameters and their influence was
provided by the study of a bench-scale planar failure, which is labelled as
case study 1. This relatively simple example of instability was analysed
using a limit equilibrium approach; providing a solution against which
numerical parameters within ELFEN, could again be calibrated. The
influence of groundwater was then investigated, highlighting sensitivities
within the comprehensive numerical approach, which were not apparent
during the analytical limit equilibrium analysis.
Chapter 5 – This Chapter introduces the second and third case study slopes, both
within a bench-scale, but involving a more complex failure mechanism.
Case study slope 2 is based upon a failure that occurred within a chalk cliff;
subsequent back-analysis of mass strength proved that the preliminary limit
equilibrium approaches provided an approximate analysis. In particular, the
slice-based limit equilibrium method gave a speculative view of mass
strength, whilst the two discrete analytical approaches demonstrated a
slightly more conservative assessment. The subsequent use of more
10
CHAPTER 1: Introduction

advanced numerical approaches proved that mass strength was in-between


the two previous assessments, with stress monitoring at several points
within the more complex ELFEN model. Finally, the development of case
study slope 3 is outlined, with the subsequent use as the first “wet”
fracturing model, in which the groundwater module within ELFEN was
tested.
Chapter 6 – Presents modelling of the failure at Delabole Slate Quarry, Cornwall.
Primarily the successive development of a model is detailed, where the
failure mechanism is accurately represented. Subsequently groundwater is
implemented firstly within a simple model, with comparison against a limit
state analysis, to calibrate numerical parameters. Finally, a rock-bridge is
introduced, which makes the model more unstable; therefore modifications
are used to decrease the influence of kinetic energy within the model,
avoiding failure of the rock-bridge during the certain modelling stages.
Consequently a model is presented which fails as a result of a rise in pore
pressure, finding the height to which groundwater must rise.
Chapter 7 – Introduces a failure at the Fimiston Open Pit in Kalgoorlie, Western
Australia. This forms the fifth and penultimate case study slope. Different
approaches are used to investigate the mass strength of a relatively small
section of basalt through which toe-breakout occurred allowing a form of
active-passive failure. Advantages and disadvantages of each approach
were examined, and a range of likely cohesive strength for the mass is
suggested. Finally recommendations are given, outlining the further detail
that is required to develop a more advanced numerical model.
Chapter 8 – The final and largest case study slope 6, is an example slope based on a
real rock mass that had been mapped and modelled for pillar strength,
during the research of Elmo (2006). This provided a source of biaxial
sections through a fractured limestone, which were used to derive mass
strength, with consideration of the orientation of applied stress relative to
dominant discontinuity orientation. Subsequently an equivalent continuum
was simulated in a large slope situation, with comparisons against
empirical-based methods.

11
CHAPTER 2: Literature Review

Chapter 2 – Literature Review

2.1 Introduction
This Chapter comprises a literature review of the factors that influence slope stability,
with Section 2.2 introducing potential failure mechanisms that can occur within a
fractured rock slope. This is followed by a review, within Section 2.3, of current
methods of numerical modelling and the division of disciplines/techniques on which
analyses are based. Subsequently several numerical modelling programs are
introduced, in order of increasing complexity. Within Section 2.4, the influence of
groundwater on slope stability is discussed, giving important issues to consider when
constructing a comprehensive fracture-based numerical model with a phreatic surface.

Section 2.5 considers the larger scheme of numerical modelling and its use to assess
rock mass strength. Primarily there is an introduction of the current empirical-based
methods by which rock mass strength may be determined, giving detail on one
particularly widely used approach. The final part of Section 2.5 reviews two new,
numerically-derived, equivalent continuum approaches to mass strength determination.

2.2 Slope Stability


Any analysis of slope stability requires: an appreciation of what may have caused or
could cause the failure, factors influencing rock strength, and the potential mechanisms
by which failure may occur.

2.2.1 Triggers for rock slope failure


For a rock slope failure to suddenly occur, there must be an anthropogenic or natural
change in the stress state, as a result of external and internal factors that act to
generate an over-stressed state within the intact rock or upon pre-existing
discontinuities. Broadly these factors can fall into one of the following areas:

 An additional load on the slope (including groundwater and dynamic loading);


 Change to the geometry (mining or erosion); or
 Internal rock mass degradation/damage.

12
CHAPTER 2: Literature Review

Rock mass damage


Rock strength can decrease over time via damage and weathering processes, this can
culminate in progressive failure of a rock slope, as discussed in Section 2.2.5.
‘Damage’ is a form of permanent strain which is endured by a plastic material (Lightfoot
and Maccelari, 1999). The potential types of damage and mechanisms by which they
occur, are listed within Table 2.1.

Table 2.1: Types of damage that can occur within slope masses, with the potential to ultimately lead
to slope failure (modified after Stead et al., 2004).

Type of damage Process


Ductile Initiation, comminution
Creep Rigid block overlying soft rock
Low/high frequency, loading/unloading (glaciation, isostasy, pore water
Fatigue
pressure fluctuation, freeze-thaw, wet/dry, seismic activity
Tectonic Preconditioned damage due to faulting, folding, in-situ stress, etc.
Anthropogenic Stress-induced damage due to excavation
Physico-chemical Hydrothermal alteration, weathering, corrosion
Stress-induced damage associated with valley formation, cambering,
Geomorphic
erosion, thermal cycling etc.

The characterisation of damage is considered essential by Stead et al. (2004), for the
successful simulation of fracture mechanics within rock slopes. However, due to
complexity, the characterisation of damage processes is difficult and only a few
comprehensive numerical codes are able to simulate certain aspects of the processes
involved.

Internal deformation mechanisms and dilation act to damage the rock mass during the
pre-failure stage upon a slope, from which tension cracks can appear; this is a form of
progressive failure, as discussed in Section 2.2.5. These mechanisms can be
particularly important in slope failures where kinematic release is inhibited, instead a
brittle to ductile transition has to occur where the rock mass strength degrades and
localised slip occurs along discontinuities (Eberhardt et al., 2004).

2.2.2 Factors that control rock slope failure


Slope failures usually involve both discontinuous and continuous aspects of the rock
mass. Failure can be upon existing discontinuities (discontinuous feature of a rock
mass) and through the intact rock failing in a brittle manner and deforming as a
continuum (Eberhardt et al., 2004). In order to understand slope failures, the factors
that control instability and determine the rock strength need to be considered. With

13
CHAPTER 2: Literature Review

many coupled processes controlling slope instability, a logical order needs to be


followed in order to build an understanding of the rock mass:

1. Attempt to understand the behaviour of a rock mass at the small-scale of


fracture mechanics;
2. Following this, build an understanding of how pre-existing discontinuities
influence rock mass strength.
3. And finally, these should be combined with consideration of in-situ stresses to
understand the development of failure surfaces during excavation (over long or
short timescales).

The following sections review each of these aspects.

Fracture Mechanics
Most rock masses are under some form of compressive stress, normally in the form of
gravitational or tectonic stress; although localised failure is often tensile. This is
because the ultimate tensile strength of a rock is much lower than compressive
strength.

All materials, and in particular rocks, have microcracks or defects often referred to as
Griffith cracks/flaws which are statistically distributed in a unit volume of a material.
These randomly orientated sub-microscopic flaws or microcracks (elements of
heterogeneity), exist inherently in brittle solids. If these extend they ultimately lead to
failure, absorbing energy via the creation of a new crack surface (Whittaker et al.,
1992).

In the vicinity of a crack tip, microcracks initiate and propagate within a zone that is
referred to as the crack tip, non-linear process zone, or fracture process zone (FPZ).
Crack initiation is controlled by the fracture toughness of the medium, refer to
Appendix A. Whittaker et al. (1992) suggests that fracture propagation is controlled by
the energy released per unit crack surface, (i.e. the energy release rate, G); thus crack
propagation occurs when the energy release rate reaches a critical value. Within
ELFEN the fracture energy parameter (Gf) is included to allow the release of potential
energy into the FPZ, and the aspects of the distribution of microcracks to be
considered within simulations (Klerck, 2000).

During fracture propagation displacement occurs at the fracture tip. Fracture mechanic
studies have identified three types of displacement, mode I, mode II and mode III.
These are briefly discussed within Section 2.2.4.

14
CHAPTER 2: Literature Review

A fracture mechanics approach to slope stability is based mainly upon the stress
intensity factor and the fracture toughness, with the ability to predict the formation of
failure surfaces within jointed rock masses and provide an explanation to how stepped
failure occurs (Whittaker et al., 1992).

Pre-existing discontinuities and rock slope mass strength


The term ‘discontinuity’ encompasses several primary (pre-deformational layering and
planes) and secondary (develop as a result of deformation) geological structures. The
less persistent primary and secondary geological fabrics and structures control the
discontinuous nature of a rock mass at small scale; whereas on a slightly larger scale it
is the interaction of the more persistent geological structures that controls the failure
mechanism. It is these more persistent geological structures (discontinuities) that can
be modelled within hybrid numerical packages. Therefore within this thesis, the term
‘discontinuity’ includes the following geological structures:

 Bedding planes (primary/pre-deformational structure)


 Joints (secondary/deformational structure)
 Faults (secondary/deformational structure)

Pine (1986) lists some of many factors that have the potential to either cause or
influence the development of existing discontinuities within a rock mass:

 Deposition
 Deformation
 Rheological behaviour
 Tectonic history
 Hydraulic conditions
 Erosion

Erosion and the subsequent elastic response to unloading, has the potential to form
many discontinuities in near-surface low stress conditions, and also open primary
features; for instance sedimentary features (cross-bedding, cleavage) can form joint
surfaces where they are exposed to weathering. As a result of unloading, sheeting and
tensile joints develop over time. Hencher and Knipe (2007) briefly discuss the potential
of a build up of water pressures within such joints, as a result of sediment infill that
restricts the permeation of groundwater out of the structure. In addition one should be
particularly aware of the development of joints in the short term; during the presentation
that Hencher and Knipe (2007) gave it was suggested that failures may occur on un-

15
CHAPTER 2: Literature Review

mapped discontinuities that have formed during the mining and construction period as
a consequence of unloading.

From the above, it is clear that discontinuities can develop over a range of time-scales;
secondary/deformational structures develop over a geological time-scale as the rock
mass comprising a slope comes into equilibrium, whereas unloading due to mining can
open joints in a relatively short time. Importantly all of these structures exist in advance
of a slope failure. During the development of a slope failure mechanism, further post-
deformational structures (fractures) may initiate. The development of post-
deformational fractures relieves stress, but causes further damage within the rock
mass, leading to the slope failure. Therefore it is important to also include fractures
within the term ‘discontinuity,’ however within this thesis, a fracture is recognised not as
a pre-existing discontinuity but instead as a feature that develops just prior to slope
failure.

Depending on the mechanism of failure that is under study, the factor of safety (FOS)
that is assigned to a slope can either refer to the strength of the rock mass or the
strength of an individual or set of distinct discontinuities controlling a potential slope
failure. In simplified terms FOS represents the ratio between resisting and driving
forces; consequently a FOS of less than 1 equates to an unstable situation whereas a
FOS of more than or equal to 1 signifies a state on the limit of stability. Importantly
FOS can be used as a design tool; in this respect a FOS it is a measure of reliability so
that one can have confidence in a slope, designed to a suitably conservative FOS.
However FOS can not be so clear-cut in complex slope situations; as discussed in
Section 2.2.5 progressive failure is a time dependent mechanism which can have
stages of movement. Therefore it can be inappropriate to assign a particular FOS in
these situations. Sullivan (2007) suggested in his presentation that slope movements
can be over large areas with failure only manifested in small areas. Therefore the
detailed characterisation of a slope failure mechanism is imperative, prior to any form
of slope stability analysis.

On a macroscopic scale the shear strength of the individual pre-existing discontinuities


and the mass strength of the intact rock, ultimately control the potential for fracture
initiation, propagation and the potential of large-scale slope failure. The rock mass
strength is primarily influenced by the geology and the relative orientation of the
structure/pre-existing fractures, which can determine the likelihood of failure via the
interaction of discontinuities. Waltham (1996) states the density, nature and extent of

16
CHAPTER 2: Literature Review

discontinuities within a rock mass, largely control the rock mass strength; although the
intact rock strength, and water conditions are also elements that should be considered.

When there are no existing adversely orientated through-going discontinuities, failure


must propagate through the intact rock mass upon extensive continuous planes. In
soft rocks, where the slope face is <45˚, the failure surface tends to be circular;
whereas in harder rocks where the slope face is steeper, failure typically follows a
near-planar mechanism upon a surface that is above the critical failure plane inclination
(pc). The orientation of this plane through a dry intact rock was defined in the
following equation from Hoek and Bray (1981):

pc = ½ (f + ) [2.1]

(Where f = dip of slope face, and  = frictional strength).

As outlined above, circular-type failures occur through the mass, where there are no
singular discontinuities persistent enough to contribute to a more planar failure
mechanism. Where there are persistent discontinuities, the failure mechanism can
become explicitly controlled by the particular discontinuity. The identification of the
failure mechanism can be through stereoplots of the discontinuities mapped within a
rock face; the types of failures that can be recognised (planar, wedge and toppling
failures) are discussed within Section 2.2.5.

Eberhardt et al. (2003) state that assumptions, such as those that within Equation 2.1
and also in failures identified through stereoplots, where failure is considered to have
occurred upon extensive and continuous planes, are justified by post failure
observations. However, any assumptions from such failures, only apply to situations
where there are persistent discontinuities and/or the volume of the failed block is
relatively small (e.g. 1000’s of m3). Larger-scale failures involve the complex
interaction between existing discontinuities and the propagation of brittle fractures.

Kimber et al. (1998) used a discontinuous numerical model, UDEC (see Section 2.3.6),
to illustrate the influence of the inclination of a single set of discontinuities with a  of
40°, termed the base plane, in comparison to the base-to-height (b/h) ratio of the
blocks that comprise the rock mass. These two controls, determine the stability of a
slope and essentially the failure mechanism. In addition Mitani et al. (2004) also used
UDEC to define the limits at which flexure toppling will occur, by determining a ratio
between the vertical height of the slope to the horizontal distance between the slope
crest and a point at the top of the cliff.

17
CHAPTER 2: Literature Review

Essentially shear failure upon existing discontinuities is controlled by cohesion and


friction. Failure occurs upon incomplete discontinuities when the cohesive strength of
the rock bridges between discontinuities is exceeded; this cohesive strength is known
as ‘effective cohesion.’ In addition to effective cohesion, the stress-conditioned
shearing resistance upon discontinuities (friction) also acts to resist shear failure
(Eberhardt et al., 2004).

The influence of in-situ stresses upon existing discontinuities should also be


recognised. This is discussed within the following section.

In-situ stresses
Most rock masses are under some form of compressive stress, normally controlled by
gravitational or tectonic stress. The action of high stresses during excavation can
compromise the shear strength of discontinuities and fundamentally the strength of the
rock mass. Pine (1986) suggests two categories of theoretical in-situ stresses:

1. Response of the rock mass as a continuum in response to gravity loading;


2. Response of the rock mass to tectonic and erosional forces, as well the ability
of the rock mass to withstand stress anisotropies.

The in-situ stress within an open pit has a significant influence on the potential slope
failure mechanisms, as discussed in Section 2.2.5. Within a low stress environment,
the rock mass dilates due to relaxation following the mining of the open pit, this can
create a blocky rock mass. Read and Ogden (2006) suggest that within a low stress
open pit mine, slope stability becomes a function of the ability of individual blocks of
rock to lock together and therefore the respective discontinuity strengths. Whereas
within higher stress environments, slope stability instead becomes a function of the
cohesion (c) and friction () or an equivalent criterion used to represent mass strength
(reviewed in Section 2.5). The influence of stress on the behaviour of rock slope is
also related to the effects of scale; as discussed in Section 2.2.3 a large slope is likely
to fail by a mass-controlled mechanism, due to higher stresses. Whereas lower stress
(bench-scale) rock slopes, are likely to fail following a dominantly discontinuity-
controlled mechanism.

With sufficient magnitude, the orientation of the maximum in-situ principal stress (σ1)
within the rock mass can have an overriding control upon fracture propagation, with
fractures forming parallel to σ1. In addition the relationship of the maximum horizontal
principal stresses (σH) to the vertical stress (σV), which is defined as the ‘k ratio’ as

18
CHAPTER 2: Literature Review

shown in Equation 2.2, has an important influence on the locality of tensile stress
accumulation.

σh = k σv [2.2]

Note that the k ratio can be dependent on Poisson’s Ratio () in sedimentary strata.
When the stress state within a rock mass is from gravity loading alone (i.e. there are no
tectonic stresses involved), the k ratio can be calculated through the following equation:

k (1or 2) 
1 
[2.3]

where k1 is in-plane and k2 is out of plane, (with no tectonic stresses k1 = k2).

Stacey et al. (2003) performed numerical modelling to prove that the location of tensile
stress within a slope can indicate the in-situ stress regime. In particular, reference is
made to different k ratios causing different behaviour within the floor of an open pit.
Stacey et al. (2003) also documented how increasing k ratio and slope height causes
an increase in the extent and magnitude of zones of high extension strain; they
illustrate a relationship with slope angle, demonstrating that in most cases the zone of
extension strain is larger with lower slope angles.

Field stresses can be altered due to excavations. It is therefore important to consider


resulting stress redistribution and create the correct excavation shape to attempt to
alleviate any unnecessary stresses. The influence of excavation shape (slope
curvature) in an open pit is reviewed within the following section.

The influence of curvature on slope stability


In the case of open pit design, the curvature of a slope can have a significant influence
on the stability. Due to the nature of stability analysis, it is often that two-dimensional
analyses (2D) are carried out over three-dimensional (3D) simulations; the benefits of
each are discussed in Section 2.3.9. Importantly, 2D models cannot consider slope
curvature. Hoek and Bray (1981) give some guidance suggesting that in some
situations slopes can be up to 10° steeper when concave, and 10° flatter on convex
slopes (when the radius of the curvature is less than the slope height), than would
otherwise be suggested using ‘conventional stability analysis.’

The beneficial effect of concave and convex slope curvature is particularly applicable in
massive rock (continuum)/non-structurally controlled slopes. Within slopes that are of

19
CHAPTER 2: Literature Review

a discontinuous nature, the beneficial effects of slope curvature are counteracted by


the control that discontinuities have upon respective failure mechanisms.

Related to slope curvature, the opposite sides of an open pit mine can interact and
influence the stability of a slope, this is a result of stress redistribution occurring when
pits are narrow. Stacey et al. (2003) note that the interaction of the opposite sides of
an open pit, occurs until the floor width exceeds about 0.8 times the slope height.
When the open pit is wider than 0.8 times the slope height, the disturbed stress field
induced from each of the pit walls is independent, therefore each slope acts separately
and consequently a smaller-scale numerical model can be used. Within narrower open
pits the opposite pit wall needs to be considered during analysis, which in some cases
requires a three dimensional model, as discussed later in Section 2.3.9.

2.2.3 Effects of Scale


There is a significant variation in rock strength depending on the scale. Sjöberg (1999)
presented a good schematic image, illustrating the relationship between strength and
volume of a rock mass, as presented in Figure 2.1a. This shows the strength of the
rock decreasing exponentially from a small and strong intact (sparsely fractured)
specimen, to the relatively intensely fractured large slope. Points upon this exponential
curve can be used to demonstrate the effects of scale, by using two slope models with
the same fracture network. One model considers the rock mass strength at bench-
scale (Figure 2.1b) and the other is to study mass strength at large-slope scale (Figure
2.1c). This also demonstrates that there is an important relationship between the scale
of study and the potential failure mechanism.

The potential failure mechanism can be related to scale-of study, as illustrated within
two example slopes which have the same fracture network but different heights,
presented in Figures 2.2b and 2.2c. Clearly defined structures are likely to
characterise a dominantly discontinuity-controlled mechanism during a bench-scale
failure (discussed in Section 2.2.5). At an overall slope-scale illustrated in Figure 2.1c,
the same fracture network instead could have a more complex behaviour and
consequent less distinct failure mechanism. Note that in this case, a circular failure
has been schematically drawn through the fractured mass. Large through-going
discontinuities have to be considered within a slope of this scale; the reader is referred
to Sjöberg (1999) for a more detailed review of potential failure mechanisms within
large rock slopes.

20
CHAPTER 2: Literature Review

Figure 2.1: Effects of scale on rock strength (a), demonstrating that a potential failure mechanism is
related to the scale of study (modified after Sjöberg, 1999).

As depicted in Figure 2.1b, depending on the network of discontinuities within the rock
mass, failures at bench-scale can occur on single a set of discontinuities if the
respective discontinuities are sufficiently persistent, being at a length that is
approximately equal to the bench height. Thus depending on the length of
discontinuity, the stability of the bench and strength of the localised rock mass can be
largely controlled by the shear strength of a single discontinuity. The shear strength
upon a single discontinuity is influenced by a number of morphological and physical
properties. Some of these properties are given by Wyllie and Mah (2004), for
instance: shape, roughness, strength of rock adjacent to the discontinuity, and
respective strength of material infilling joint. In addition to this the groundwater
conditions upon the joint also dictate the shear resistance; all of these parameters are
integrated into the empirical classification based methods of rock mass strength
determination, and should importantly be essential components within any field data
collection to allow more detailed numerical modelling.

On the smallest scale between the spacing of individual joints, failure through intact
rock is controlled by the fracture mechanics of the medium. Such a scale is important
to the understanding of fracture propagation, although this research should consider
the strength of the rock mass (which includes consideration of joint strength). Wyllie
and Mah (2004) list the following factors encompassing both the control of the intact
rock and the strength of discontinuities, which should be considered in order to
describe the strength of a jointed rock mass:

21
CHAPTER 2: Literature Review

 Compressive strength and friction angle of the intact rock;


 Spacing of discontinuities and the conditions of their surfaces.

In addition to these, other properties of the discontinuities (such as their orientation and
persistence), and also other characteristics related to the material comprising the rock
slope (such as its fracture toughness, tensile strength, etc.), ultimately influence the
strength of the rock mass.

The derivation and subsequent scaling of parameters becomes an issue when


considering the strength of a rock mass. Laboratory testing of discontinuities is usually
on a very small scale in comparison to the size of the field-scale discontinuity upon
which deformation is being studied. Curran and Ofoegbu (1993) state that the scaling
of data from laboratory tests, to represent the mechanical behaviour of discontinuities
in situ, is a crucial and unresolved problem. They also suggest that when scale
increases, the difference between the peak and residual strength becomes minimal.
This is because when scale increases the strength of a discontinuous mass decreases
considerably, as illustrated in Figure 2.1a. Consequently there is a movement towards
more mass/residual behaviour, thus the difference between peak and residual strength
diminishes, although importantly there is still a distinction.

Wyllie and Mah (2004) describe two frequently used empirical methods that have been
derived and developed as a result of the difficulty in relating an estimate of shear
strength of the rock mass, to the shear strength of relatively small samples. These two
methods are ‘back analysis’ and the ‘Hoek-Brown strength criterion;’ both of these
allow the derivation of rock mass strength, and are independently discussed in
Section 2.5.

The back analysis of a slope failure requires detailed characterisation, concerning the
conditions within the rock mass immediately before failure, thus developing an
understanding of the conditions that in theory were on the limit of stability equilibrium
(FOS = 1; as discussed in Section 2.2.2). This information could be the position of the
sliding surface, the ground water conditions, and detail on the existence of any external
forces, foundation loads, earthquake forces etc. (Wyllie and Mah, 2004). With this
information, analyses can be performed with trial and error of the likely rock mass
and/or discontinuity strength, until a condition is achieved where the slope is at its limit
of stability. Some modern numerical analysis programs can perform automated
detection of the limit state, such as Phase2 (Rocscience, 2008), through the application
of the ‘shear strength reduction’ (SSR) technique (see Section 2.3.4).

22
CHAPTER 2: Literature Review

2.2.4 Modes of fracturing


At a small-scale, three fundamental modes of displacement can occur at a crack tip; all
of these ultimately result in fracture propagation and thus can be referred to as modes
of fracture propagation. Each mode of fracture has a distinct relative movement
between fracture surfaces at the time of fracture initiation, as illustrated in Figure 2.2:

 mode I: normal stress causes tensile forces, resulting in extension/opening of


the crack tip and consequent movement perpendicular to the fracture surfaces
(crack faces move apart symmetrically);
 mode II: in-plane shear stress causes a sliding movement (shear)
perpendicular to the fracture front and parallel to the fracture surfaces;
 mode III: anti-plane shear causes a scissoring action via movement both
parallel to the fracture front and parallel to the fracture surfaces.

Figure 2.2: Modes of fracturing (after van der Pluijm and Marshak, 1997).

In addition to the above, mixed modes of fracturing can occur, which entail
displacement that is both parallel and perpendicular to the fracture surface. The
fracturing of rock is often via both opening and sliding displacements causing mixed
mode I-II fracture (Whittaker et al., 1992).

At present the solution procedure within ELFEN has the ability to model fracture
propagation via only mode I fracturing. This could limit the ability of ELFEN to predict
fracture extension within some slope problems. However, the fact that current version
of ELFEN can only simulate fracture extension though only mode I behaviour is not as
critical as it may seem. Although the common stress field in a rock mass is

23
CHAPTER 2: Literature Review

compressive, extensive tensile microfracturing is often the precursor, with shear


fracture being the final stage (Mas Ivars et al., 2007).

The application of the “combined Mohr Coulomb with Rankine tensile cut-off” criteria,
(see Appendix B), allows the consideration of tensile (mode I) and shear (mode II)
failure. Through research within this thesis, the combined criterion has permitted the
analysis of peak rock mass strength, accurately simulating mass-controlled failure up to
yield point. Following yield point, failure mechanisms involved with mass-controlled
slope failure can only, using the present version of ELFEN, be simulated where there is
a degree of discontinuity control, and orientation is such that tensile fracturing will
cause kinematic release.

It could be considered that shear fracture extension within a compressive stress regime
(mode II failure), is essential to achieving kinematic release in many rock slope
failures; therefore this is an important aspect that needs to be addressed in code
development. In the meantime it must be recognised that using the current version of
ELFEN, the simulation of post-yield failure mechanisms within rock slopes where there
is a degree of mass-control, can only occur within some types of toppling failure and
certain step-path failure situations; both of these are discussed within the following
section.

2.2.5 Ultimate failure mechanisms


The categorisation of a failure mechanism is an inherently complex task; failure
mechanisms can change with time, as failures often involve a number of mechanisms.
Stacey (2007) suggested as a result of this, that multiple criteria should be considered
when modelling. In part, this is possible in ELFEN and is incorporated within models
via the coupled Mohr-Coulomb and Rankine criteria (Appendix B).

Critically one should respect that slope failure mechanism is scale-dependent, as


discussed in Section 2.2.3. Duran and Douglas (2000) suggest that there is invariably
a structural control to failures that occur within slopes with a height less than 40m.
Within this research failure slope failure at this scale (< 40m) are termed ‘bench-scale
failures.’ Four modes of slope instability are typically identifiable at this scale: plane,
wedge, toppling (direct and flexural) and circular; these are firstly introduced within this
section. Each of these involves the interaction of existing discontinuities and can thus
be identified through kinematic analysis using stereoplots.

24
CHAPTER 2: Literature Review

At rock slopes that exceed 40m in height, large-scale plane and wedge failures
become uncommon, as a result of few persistent discontinuities that reach this scale.
Instead at large slope scale, the interconnection of discontinuities (step-path failure)
becomes important, to create rotational shear (a form of circular failure), toppling and
active-passive wedge failure.

Firstly, bench-scale failure mechanisms are discussed below, followed by an


introduction to the active-passive wedge failure mechanism. Finally there is a detailed
outline of progressive-type rock slope failures, with discussion of the important
processes that should be considered when modelling failures with significant
progressive-type behaviour.

Bench-scale failure
Large-scale planar failure is thought by Wyllie and Mah (2004) to be rare. As
discussed above, planar failure is a mechanism that instead is relevant to slope
instability at bench-scale. The conditions for a planar failure rely upon the existence of
a slide plane that daylights into the slope surface, striking parallel or nearly parallel to
the slope face. A tension crack behind the slope crest, along with release surfaces at
either end of the failure plane, enables kinematic release of the failure block. Within
this thesis, a planar failure has proved valuable for demonstrating the sensitivity a
comprehensive stability solution has, to variation in surface properties; this is outlined
in Chapter 4.

Wedge failure is concerned with two intersecting discontinuities which strike oblique to
the slope face. At bench-scale, the majority of discontinuities are often small in length;
therefore for wedge failure to occur, two unfavourably orientated discontinuities must
exist. Considering that rock masses have few large discontinuities, it is unlikely that
orientations will be such that large wedge failures will form. Therefore this failure
mechanism is limited to relatively small-scale failures affecting a limited area; the shear
strength upon the discontinuities provides the overriding control on failure. At bench-
scale, it can be considered a more probable failure mechanism than planar failure, due
to the wide range of geologic and geometric properties with which wedge failures can
occur (Wyllie and Mah, 2004). The line of intersection which is formed by the two
planes ultimately controls the stability, with failure occurring when this line of
intersection dips at an angle that exceeds the frictional angle of the respective
discontinuities.

Unlike planar and wedge failure, toppling failure does not involve the sliding of a block;
instead toppling failure is concerned with the rotation of columns or blocks of rock
25
CHAPTER 2: Literature Review

around a fixed base (Wyllie and Mah, 2004). Several types of toppling failure exist,
although the most important distinction is between direct and flexural toppling; both
occur when there is a set of discontinuities steeply dipping into the slope face.

Flexural toppling can occur with only a single set of discontinuities dipping steeply into
the face forming continuous chimneys of rock that can flex and consequently break,
allowing release from the slope face. Direct toppling however, requires another set of
orthogonal subhorizontal discontinuities, forming the release surfaces for blocks to
topple directly from the slope face. Mitani et al. (2004) used UDEC to demonstrate that
the initial failure of a flexural topple is caused by tensile stress between critically
dipping failure planes; in addition shear failure can accompany the flexure topple
through the interlayer slip as a result of the outward movement of each column of rock.

Finally failures within either weak material or closely/highly fractured rock masses,
follow a circular geometry, involving failure through entirely intact rock as a result of its
weak nature, or upon existing discontinuities. During circular failure, failure starts at
the toe of the slope and gradually works upwards, in a time-dependent fashion (see
following section on progressive failure), as favourably orientated fractures are linked
via the fracture of the intact material (rock-bridges) between the discontinuities. As a
result of this nature, the material through which circular failure occurs could be
considered a continuum and thus can be applicable to limit equilibrium and finite
element solution schemes introduced in Section 2.3. The geological conditions within
the slope ultimately control the shape of the failure surface, with shallow, large radius
surfaces occurring in homogeneous weathered or weak rock masses; whereas in a
more cohesive and lower friction strata (such as clay), the circular failure surface is
deeper and has a smaller radius that may daylight in an area beyond the toe of the
slope (Wyllie and Mah, 2004).

Failure in large-scale slopes


Stacey et al. (2003) suggest that large deep seated rotational failures form as a result
of zones extensional strain; consequently extensional events, which can be detected
by seismic monitoring, could mark the onset of fracture extension within large circular
failures. In addition, large rock slopes can be under considerable compressive stress,
and the progressive failure within these highly stressed zones can result in strain
bursting, as in deep underground mines. Stacey et al. (2003) suggested that slab
failure upon the slope face, and popping up of the pit floor, are physical manifestations
of strain bursting. Consequently there should be a view forward, to link models of
large-scale circular failure, to seismic monitoring data.

26
CHAPTER 2: Literature Review

However, circular failure through relatively brittle rock masses that comprise a large
slope is a controversial issue. Conventionally, the analysis of such failures has been
via continuum-type circular failure (mass-controlled) methods; however the influence of
discontinuity-control is becoming more of an important aspect with deepening open pits
creating larger slopes. Consequently the simulation step-path failure at a large-scale,
is now a key issue. This requires an increase in both computer power and reliability, in
order for accurate representation via comprehensive fracture-based numerical
modelling methods.

At present, equivalent continuum analyses (introduced within Section 2.5) using finite
element or finite difference models, are the principal way in which large-slope failure
mechanisms can be represented efficiently. Using this technique, the simulation of
discrete large-scale structure within large slopes, can only occur via accurately
assigned biaxial strength models, with consideration of the representative elementary
volume (REV) of the rock mass. The accurate simulation of progressive failure at
large-scale within a rock slope, still remains a key issue than cannot be directly
represented using an equivalent continuum technique. Despite this, the final part of
this section introduces progressive failure, with a view forward to future simulation
within large fractured rock slopes. Before this, active-passive wedge failure is
introduced within the following section.

Active-passive wedge failure


The active-passive wedge failure is a bi-planar mechanism whereby there is an upper
active block, this can be formed by an active fault; the weight of the block causes the
failure of a passive wedge that is defined by an underlying low angle failure surface
such as intra-formational shears or weak bedding planes (Stead et al., 2006a). Such
failures are common within UK opencast coal mines, where the lower basal plane is a
plane of weakness such as a seatearth, with kinematic release via a rear failure
surface such as a fault (Stead and Coggan, 2006).

The kinematic constraint within active passive wedges/biplanar problems requires a


modelling approach, such as ELFEN, that allows progressive fracturing (Stead et al.,
2004). The intact fracture and discontinuum abilities of ELFEN allow the simulation of
the complex failure within active-passive geometries.

The geometry of an active-passive wedge failure can be considered very similar to


failure on a planar discontinuity, where kinematic release is from a tension-crack. The
only difference between these two geometries is that the tension crack within the
27
CHAPTER 2: Literature Review

planar failure is vertical, whereas with an active-passive geometry the release surface
is non-vertical; the consequence of this is a ‘chair-shaped’ failure scar.

Stead et al. (2006a) modelled an active-passive problem which demonstrated how the
fracturing of the upper and/or lower blocks, provides space for kinematic release. The
nature of kinematic release within an active-passive wedge failure is time dependent;
the processes controlling time dependent deformation are discussed within the
following section.

Progressive failure
Progressive failure is time-dependent deformation that allows the eventual completion
of a failure plane via mechanisms (stress corrosion/damage and chemical/physiological
changes), which progressively reduce the strength of discontinuities and ultimately the
rock mass. Such a process can lead to one of many failure mechanisms resulting in
slope failure in either a brittle or ductile manner.

Progressive failure is an especially important failure mechanism at a large-scale; as


discussed in Chapter 1, large slopes are becoming a more frequent global mining
practice, and therefore the understanding of progressive failure can be considered vital.
Franz et al. (2007) suggests that step path failures are particularly important within
large scale slope failures where often, failure is across several separate discontinuities
that exceed bench heights.

Stead et al. (2004) described a step-path failure. With associated time dependency,
step-path failure can also be considered as a form of progressive failure. The
generation of a step-path failure is greatly influenced by the persistence of a low angle
joint. As illustrated in Figure 2.3, the geometry that Stead et al. (2004) presented
clearly shows persistent low angle joints, which promote the extension of cross-over
fractures and allow subsequent release of the discrete block. If joints are not persistent
enough, the overlap may be insufficient for a cross-over fracture to form through the
rock-bridge. Yan et al. (2007a) describe numerical modelling that has been used to
simulate such factors, which are important in step-path failure. Also, as referred to
later within this section, Yan et al. (2007b) suggested a number of properties which
determine the failure potential of a rock-bridge, one of these is the persistence of the
step-path fractures.

28
CHAPTER 2: Literature Review

Figure 2.3: Step-path failure showing persistent low angle joints (after Stead
et al., 2004).

This propagation of the cross-over fractures to the kinematic release of the failure
mass, characterises how progressive failure involves a change in kinematics at micro
to macro-scales. As the shear strength along the failure plane reduces from peak to
residual values, consecutive failure of rock bridges can lead to the development of a
failure surface upon which kinematic release may become possible (Eberhardt et al.,
2004).

The following sections review this mechanism of failure, with examples where
numerical models have been used to simulate behaviour and consequent suggestions
upon inadequacies within the present modelling approaches.

2.2.6 Behaviour of a Progressive Failure


A progressive failure mechanism represents time-dependent deformation of a rock
mass; consequently different stages can be identified during the failure process.

Sullivan (2007) suggests that most open pit slopes pass through five stages of
successive movement:

1. Viscoelastic response.
2. Creep.
3. Cracking and dislocation.
4. Collapse or failure.
5. Post failure deformation (“liquefaction type failures”).

The first four of these stages of movement are illustrated in Figure 2.4. Primarily the
initial response of the rock mass is characterised by a sudden rapid increase in the
deformation rate, which can be associated with the primary stage of movement
illustrated in Figure 2.4. However it is from the secondary phase of deformation, that

29
CHAPTER 2: Literature Review

behaviour differs between stable pit slopes and progressive-type failures; Mercer
(2007) examines the behaviour of a progressive failure in detail, reporting a rapid
decrease in deformation rate, following the initial stage. Also finally there is a long
period of slowly reducing steady state creep.

Figure 2.4: Stages of movement that occurs within moving slopes,


identified by Fell et al. (2000, cited in Sullivan, 2007). The early
phases of deformation can be related to the style of deformation that
can occur during progressive failure.

Subsequently Mercer (2007) categorises five distinct stages of deformation within a


progressive failure:

1. Primary rock mass creep during a pre-collapse stage;


2. Secondary rock mass creep during a pre-collapse stage;
3. “Post-onset-of-failure to collapse behaviour mode;”
4. “Post-collapse behaviour mode;”
5. “Post-mining/recovery behaviour mode.”

In addition Mercer (2007) outlines how the post-deformational behaviour of a


progressive failure can be complex, with the potential of alternating progressive and/or
regressive behaviour.

The two most important areas that should be considered when studying a progressive
failure is the strength of rock bridges between unfavourably aligned discontinuities, and
the time dependent development of fracture through these rock-bridges ultimately
creating the failure path. Yan et al. (2007b) suggest that the failure of rock bridges is a
function of persistence, dip, joint spacing, shear strength and the intact tensile strength;

30
CHAPTER 2: Literature Review

Elmo et al. (2008) look towards the characterisation of rock-bridges suggesting that
fracture spacing, persistence and block size can define a rock-bridge. It can be
anticipated that there is a direct relationship between the size of rock-bridges and the
resulting rock mass strength; the larger the rock-bridge the less likely fracture will be
able to span the distance between the respective discontinuities. As discussed in
Section 2.5.2, Elmo et al. (2007b) introduces a relationship between fracture intensity
and the spacing and size of rock-bridges, and subsequent rock mass integrity. The
only disadvantage to this approach is that their ratio cannot be used as an index of the
physical dimensions of the rock bridge; instead Elmo et al. (2007b) suggests that this
should be derived from an interlinked rock mass classification system.

Time dependent development of fracture through rock-bridges is a problematic issue to


address due to the many variables that need to be considered. Starfield and Cundall
(1990) identified that a time-dependent constitutive model is needed for fractured hard
rock. However, as many authors have suggested, Harrison and Hudson (2003) state
that there is still insufficient knowledge of time dependent processes in fractured rock
masses. Chung (2007) recommends that the time dependent behaviour and long-term
stability of a rock slope can only be better understood by the characterisation of shear
creep within rock masses. In order to build an understanding of this, the processes
causing and driving progressive failure need to be understood.

2.2.7 Triggers and processes driving progressive-type failure


As discussed earlier, an open pit slope may inherently exhibit a form of progressive-
type deformation; mining excavation is therefore an important and common trigger
mechanism. Eberhardt et al. (2004) also suggests the following triggers for
progressive failure:

 Gravitational loading of a continuum, resulting in excessive shear stresses and


strains;
 Delayed response of rock mass towards glacial processes (in some areas this
can be considered the most recent geological process, which would have had
caused a change of equilibrium conditions).

Once the progressive behaviour has been triggered, and it is following the style of
deformation outlined earlier within this section, one must consider the underlying
processes that can control and drive progressive-type behaviour. Eberhardt et al.
(2004) outline how the process of progressive failure is driven by:

31
CHAPTER 2: Literature Review

 The propagation of fractures (breaking intact rock bridges that exist between
discontinuities), thus overcoming the cohesive strength of a shear plane;
 Strength degradation, in particular the frictional strength is reduced as a result
of mobilisation causing the shearing of asperities along the respective
discontinuity surfaces;
 Strain softening; and
 Internal deformation accompanied by dilation causing such failure events to last
several hours, with the tilting and falling of large blocks one after another; as
was reported within the events of the Randa 1991 slide.

The degradation of strength within a rock mass was briefly discussed in Section 2.2.1.
In more detail Hencher and Knipe (2007) introduced three ways via which progressive
failure and a consequent ‘dynamic evolution’ of the strength may be induced:

 “Changes in the cohesion between mineral phases across grain boundaries”


 Slow mineral dissolution causing an expansion in porosity, or chemical reaction
to weaker phases;
 The removal of micro-bridges which were previously load-bearing.

In addition to some of the above Wyllie and Mah (2004) suggest that increases in pore
pressures over time (including swelling) and time-dependent creep as a result of a
constant load, are also both mechanisms of progressive failure. Analogous to these
processes are the forms of damage, outlined previously in Table 2.1, which can also
promote the progressive failure of rock slopes. One of these forms of damage is
physico-chemical; Mas Ivars et al. (2007) refers to stress-corrosion (chemically
assisted crack growth) being responsible for time-dependency behaviour during brittle
creep. They also noted that a Particle Flow Code, which is introduced in Section 2.3,
possesses a ‘parallel-bonded’ stress corrosion model, which is able to simulate this
time-dependent behaviour.

Progressive failure can be instigated by excessive stresses; the point at which this
stress is relieved could manifest itself in either internal or external deformation,
therefore the factor of safety drops below unity at this point. However, following stress
relief the failure mass can enter a semi-stable period during which shear stresses build
until they are sufficient enough to cause failure again. This sort of stick-slip behaviour
cannot be assessed by conventional approaches which give a single FOS; also time
dependent behaviour, is a difficult aspect to represent when modelling. Different
mechanisms can be involved in the initiation and development of a progressive failure;
whilst a large and deep seated mass can be involved at the initiation stage, the
32
CHAPTER 2: Literature Review

consequent collapse can occur within a relatively small zone near surface (Stacey,
2007).

In-situ stress is an important factor that can control progressive-type failures, it must be
noted that progressive and step-path failure can occur both in low and high stress
situations, although differences exist between the mechanics during the step-over
fracture through rock-bridges. In low stress environments it has been shown that
failure within intact rock bridges is often tensile even if the surrounding stresses are
compressive; under a higher confining stress shear failure of rock bridges becomes
more important (Lajtai, 1969, cited in Alzo'ubi et al., 2007). Therefore in order to model
large-scale progressive-type failures, the numerical code should be able to
insert/develop fractures as a result of mode II failure (introduced in Section 2.2.4). This
is discussed further within Section 2.2.9, which outlines the capability of current
numerical modelling methods, to simulate progressive failure. Before a model can be
constructed, one needs to consider examples of progressive failure and how these are
monitored.

2.2.8 Examples of progressive failures


As previously discussed, most slopes show progressive-type failure; monitoring can
show the development of this behaviour within some slopes. For instance Harries et al.
(2006) analysed a multi-bench day-lighted circular (slump) failure at Mount Owen Coal
Mine, Australia. Graphs of displacement were presented at certain points within the
moving mass showing a characteristic acceleration; Harries et al. (2006) concluded
that this accelerating behaviour may mark the subsequent progressive development of
the circular failure surface, until kinematic release.

Another form of progressive failure involving the degradation of slope was suggested
by Roux et al. (2006) in their assessment of the 1997 failure at Navachab Gold Mine,
Namibia. They suggested that the failure was primarily a result of a reduction in the
shear strength on critical joint planes, due to both blasting and changes in the in-situ
stress state during mining which caused a state of over-stress and an overall
weakening of the rock mass. Another example where monitoring is very important, is
given by Kveldsvik et al. (2006), who describe a secondary or steady creep phase of
deformation in the Norway Fjords. Monitoring slope deformation therefore allows the
identification of critical acceleration periods during progressive-type behaviour, which
occur prior to a catastrophic failure.

33
CHAPTER 2: Literature Review

The sudden/rapid failures on unknown structures, reported by Dight (2006) and


Simmons and Simpson (2007), are likely to be forms of progressive failure. Although
there has not been any direct reference to progressive-type behaviour, they involve a
degree of brittle failure through the rock mass, which itself is inherently time-
dependent.

An important example of progressive behaviour is presented in Eberhardt et al. (2004),


who modelled the Randa 1991 slide. The processes involved with progressive failure
can be used to explain observations during slope failure. During the Randa 1991 slide,
direct observations were made of the time-dependent process. Eberhardt et al. (2004)
suggest that the Randa 1991 failure was relatively slow as a result of internal
fracturing, which also caused the dissipation of energy and the subsequent short
runout of debris. These pre and post-failure conditions should be noted and
considered as features that may indicate progressive-type behaviour. The following
section gives some of the more important features that have to be addressed when
modelling progressive failure.

2.2.9 Numerical modelling of progressive failure


Within progressive failure it is clear that there is a definite transition between brittle and
ductile behaviour. Sainsbury et al. (2007) suggest that the specification of brittle or
ductile behaviour within a numerical model is very important with regard to progressive
failure, which is likely to occur much sooner after yielding within brittle materials rather
than in ductile materials; consequently they used a strain-softening model to represent
the post-peak strength degradation which accompanies failure. Tang et al. (1998)
suggest that RFPA2D (Rock Failure Process Analysis - which is a linear continuum
programme that simulates rock fracture), can model progressive failure that often leads
to collapse in brittle rocks.

The ability of ELFEN to model the processes involved with progressive failure is
ensured through the remeshing and contact search algorithms within the coding,
although perhaps ELFEN is limited to simulating fracture during progressive behaviour,
in only low stress environments. As discussed previously, step-path failure in high
stress situations is likely to involve mode II failure; as discussed in Section 2.2.4,
ELFEN cannot insert a fracture when failure of the rock mass is by mode II failure.
This may limit the simulation of progressive failure in large rock slopes where stresses
are considerable.

34
CHAPTER 2: Literature Review

Importantly, conventional rock slope analysis methods cannot account directly for the
progressive development of a failure plane. Instead most methods only address joint
persistence; Eberhardt et al., (2004) suggest that the consideration of joint persistence
in conventional analyses, involves limit equilibrium solutions that apply an apparent
cohesion, which is dependent on the continuity of jointing. The tensile fracture ability of
ELFEN enables the direct modelling of step-path and consequent progressive failure.
During the step-path simulations by Yan et al. (2007a), a linear post-yield strain-
softening response is assumed following points at which the tensile stress reaches the
tensile strength. The gradient of the linear strain-softening response curve is
calculated through ELFEN and is defined by a formula, based mainly upon the fracture
energy (Gf), presented within Klerck et al. (2004). There was a brief introduction of the
Gf parameter in ELFEN, within Section 2.2.2; further discussion is provided in
Appendix A.

Aspects of progressive failure have also been modelled in UDEC (introduced in Section
2.3.6); Alzo'ubi et al. (2007) recreated some of Lajtai’s (1969) experiments using their
UDEC damage model (UDEC-DM). They concluded that the Voronoi method within
UDEC, is capable of accurately simulating discontinuous behaviour of fractures in
direct shear.

At present most numerical analysis of progressive failures within rock slopes, have
been based in two-dimensions (2D); however as discussed in Section 2.3.9, some
situations require analysis in three-dimensions (3D). Read (2007) suggested that the
simulation of step path failure can be restricted in 2D models due to the formation of
discrete blocks or closed volumes from the extension of intersecting structures. This is
perhaps less likely in 3D analysis. Further research is required to verify these findings.

Yan et al. (2007b) used 3DEC (introduced in Section 2.3), to simulate translation and
rotation of rigid blocks, subsequently forming a step-path failure surface within a slope
where joints were persistent. ELFEN was also used by Yan et al. (2007b), for the
simulation of less persistent fracture networks, when modelling the failure of rock
bridges.

As discussed in Section 2.5 and in Chapter 8, typical modelling of fractured rock


masses, is to simplify them via an equivalent continuum approach. Ubiquitous joint
models and anisotropic rock mass models are methods by which one can consider
step-path failures within advanced simulations of an equivalent continuum; however
results from such analyses can often be conservative and even misleading (Franz et

35
CHAPTER 2: Literature Review

al., 2007). The findings of Elmo et al. (2008) are also significant; reporting that both
equivalent continuum approaches and rock mass classification schemes can result in
an underestimation of the rock mass strength properties. This could be because rock-
bridge strength cannot be considered within these conventional approaches.

Therefore accurate modelling of a progressive failure is only achievable through direct


modelling of the discrete features of a rock mass; however as discussed in Sections
2.3 and 2.5 this is not always possible due to limited computational power.
Consequently considering the present computational capability, equivalent continuum-
based approaches to modelling progressive failure within rock slopes have to be
accepted, but with the view that mass strength estimation is likely to be conservative.

For the purpose of this thesis mechanical issues related to failure within fractured rock
slopes have been addressed. During the review of these processes, it has been made
clear that there are human-based factors (for example the amount of empirical
knowledge during a design situation), which can act to compromise slope stability;
some of these issues are discussed briefly within Appendix D.

It appears imperative that some of the issues discussed in Appendix D, are addressed
if large slopes are to be designed with confidence. With the increasing use of
numerical models worldwide by both consultancies and mining companies, it is a
priority that a detailed understanding of numerical modelling is gained during university
teaching. The following section reviews in detail the numerical models that are
available for fractured rock slope modelling, and the concepts upon which they are
based.

36
CHAPTER 2: Literature Review

2.3 Methods of Numerical Modelling


Numerical modelling is a method of arriving at an approximate solution for stress
distribution and displacements surrounding an excavation (Meyer, 2002). Numerical
models cannot be an exact imitation of reality; instead they should provide a simplified
representation of the complex real situation. Starfield and Cundall (1988) particularly
stress the ability of numerical models to reveal the processes that can otherwise be
overlooked. One of their main suggestions is that: “A model is an aid to thought, rather
than a substitute for thinking” and consequently they propose a set of guidelines that
should be followed to form a methodology, prior to the modelling exercise.

Often models may simply be used to investigate theories on the process of failure
initiation; another objective could be that modelling rock masses gives an insight into
which processes should be considered explicitly in an average way (Hoek et al., 1990).
The creation and formulation of a numerical model is invariably conducted via
computer software due to the sheer complexity and number of calculations required
during computation. Such models make it possible to investigate a number of
alternatives in a realistic manner, but it should always be referred to as only an
approximation to reality (Hoek, 2004).

Siddall and Gale (1992) suggest that the use of any computer simulation is reliant
primarily on the data used, the skills and knowledge of the modeller. Simulations
should therefore be used as an assessment of data gathered from monitoring
programs, allowing a better prediction of the likely response of strata to various mining
situations, than any empirical method can deliver. Therefore there are distinct
advantages of numerical modelling over empirical techniques; however the
complicated behaviour of a rock mass cannot be captured by even the most
comprehensive numerical modelling packages. Consequently the use of empirical and
numerical techniques alongside each other should always be considered in the future
development of predictive techniques for rock mass behaviour.

37
CHAPTER 2: Literature Review

2.3.1 Overview of types of numerical modelling software


Numerical models range in their complexity, each suited to particular applications; from
the relatively simple infinite slope and planar failure limit equilibrium methods to the
more complex coupled finite-/distinct-element codes (Eberhardt, 2003). Stead et al.
(2006a) introduces the concept that three ‘levels’ of sophistication can be considered
within rock slope analysis;

I. Conventional kinematic and limit equilibrium techniques


II. Continuum and discontinuum approaches
III. Hybrid methods

The research in this thesis uses all three levels; although the majority of the work has
been completed using level II and III methods, with the use of limit equilibrium (level I)
analysis as a check on the behaviour of ELFEN, which falls into the third level of
numerical analysis. The following Table (2.2) gives an overview of numerical
techniques that are currently available; outlining the basis of each the method of
analysis, its appropriate area of application within the geotechnical study of slopes, and
finally the benefits and limitations of each method. The information within the table has
been composed from a collection of sources, all of which are accordingly cited.

At the simplest level of analysis, kinematic methods consider the rock mass as a series
of blocks, identifying those blocks that are unstable. Poropat and Elmouttie (2006)
present a method of performing 3D structural modelling of open pits by acquiring
structural orientation data via Sirovision 3D imaging system. Following this a block
analysis is performed to detect which blocks have the potential to be kinematically free.

Fundamental issues have to be addressed in order to accurately represent a fractured


rock slope, when using the more advanced forms of analyses (level II and III). Several
different approaches exist; in order to understand the output of a numerical model it is
essential to have an awareness of the approach upon which the computational method
is based.

A numerical model firstly either has a continuum or discontinuum foundation; both of


these contain elements of some sort, within which stresses and strains are
independently solved. The distinction and subsequent choice between these methods
is illustrated clearly by Scheldt et al. (2003) in Figure 2.5.

38
CHAPTER 2: Literature Review

In addition to the material-basis within a numerical model, there are fundamental


differences between the meshing (to create elements) and simulation procedures. The
following bullet points, briefly detail the differences between domain and boundary
numerical models (which are related to the meshing procedure), as well as the
distinction between implicit and explicit models (related to solution procedure) within
level II and III methods:

 Domain or boundary
- Domain solutions discretize the whole model into elements.

- Boundary solutions only discretize boundaries within the model.


Subsequently elements are created along boundaries such as
excavation surfaces, joints and material interfaces, with the interim
material being left as an elastic continuum (Flynn, 2001).

 Implicit or explicit
- The explicit scheme is based on the fact that there is a maximum speed
at which information can propagate through elements; consequently the
time-stepping procedure is employed simplifying the calculation
procedure. This allows each element to be considered independently;
effectively the model is therefore idle whilst new properties are
calculated for each element (Flynn, 2001). It must be noted that the use
of explicit analysis (as opposed to the implicit approach) is more suited
to modelling non-linear large strain problems.

- Implicit approaches also have a time stepping solution procedure,


however elements are not considered independently, instead during
each solution step there is communication between elements;
consequently for an equilibrium to be reached in each solution step
often several iterations are required (Flynn, 2001).

39
CHAPTER 2: Literature Review

Figure 2.5: Decision tree for selecting modelling method (after Scheldt et al., 2003).

Given an understanding on how a numerical model is composed, one can consider the
suitability, and the consequent benefits and limitations of different numerical
approaches; as presented in Table 2.2. Following this, there is further discussion upon
the detailed basis of some of the numerical approaches introduced within Table 2.2.

40
CHAPTER 2: Literature Review

Table 2.2: Basis of some of the more frequently used methods of slope stability analyses that are available, with suitability, benefits and limitations.
Basis of model and
Composition Suitability Benefits Limitations
method of analysis
2D rigid blocks Can be quite advanced with examples such as:
 coupled groundwater simulations;
Best suited to the analysis of
 implementation of probabilistic
relatively simple trigger
Can be
mechanisms. No ability to consider complex
considered as an approaches to numerical modelling
Kinematic and behaviour such as:
 internal/intact brittle
analytic (and non (which can be used to generate a
Limit Equilibrium Most applicable to simple
numerical) factor of safety, for the purpose of back
Techniques block failures along a fracture and deformation,
 progressive failure/creep
method of analysis a problem*).
discontinuity or rock slope
stability Most commonly used techniques for the
that behaves like a continuum
assessment evaluation of stability within open pit mine
(Eberhardt, 2003).
(Chung, 2007). slopes Stacey et al. (2003).
 Requires specialised input
Capability to model:
 internal deformation within intact
parameters.

 Experienced user required.


material
 material heterogeneity,
 non-linear deformability (mainly
 Segregation of whole of study
Failures which are controlled
by deformation of the intact plasticity),
 complex boundary conditions,
Continuum domain area into elements, thus a more
rock mass within:
 in situ stresses, and
Finite elements extensive mesh necessitating
  gravity.
FEM – Finite Element greater runtimes than the BEM
Heavily weathered or
Method Uses implicit
 These greater runtimes limit the
fractured rock

solution
Weak rocks/soils Joints can be included but are represented
FDM – Finite procedure ability to run sensitivity analyses
implicitly through the ‘equivalent continuum
Difference Method (Stead et al., 2006a).
(i.e. rock mass behaves approach.’ **
 Not able to model detachment of
as a continuum).
Ability to import slope cross sections directly
elements or large rotations.
from digital terrain models.
 No appreciation of crack tip
Current development of meshless (or ‘mesh-
stress concentrations.
free’ and ‘element free’) FEM approaches.
Notes: *Back analysis of a slope failure allows the derivation of shear strength parameters for a particular rock mass.
**The equivalent continuum approach involves the weakening of mechanical properties to give an overall behaviour including incorporation of the response
41

of both joints and the intact rock (Curran and Ofoegbu, 1993).
41
CHAPTER 2: Literature Review

(Continuation of Table 2.2 - Basis of some of the more frequently used methods of analysis that are available, with suitability, benefits and limitations)

Basis of model and


Composition Suitability Benefits Limitations
method of analysis

Creation of an excavation without influencing  Requires specialised input


far-field stresses, which should remain linear- parameters
Finite elements
elastic.
 Experienced user required
Continuum boundary around only the
Applicable to large models Jing (2003) considers the BEM to be:
 most efficient in the modelling of
boundary of the
within homogeneous and
 Not able to model large rotations;
BEM – Boundary model, interior is
linearly elastic bodies fracture propagation, and
 more accurate at the same level of
Element Method an infinite
 No appreciation of crack tip
continuum
discretization than other continuum
methods stress concentrations.

 Requires specialised input


parameters, in addition joint
properties are required that are
Discontinuum not routinely measured
Has all the benefits of the FEM and FDM
domain
 Experienced user required
approaches but is also able to model:
Discrete  displacement and deformation along
DEM – Discrete
 Models can require long
elements individual explicitly modelled
Element Method
Applicable to fractured/blocky discontinuities;
 highly non-linear behaviour which occurs
runtimes
Uses explicit rock masses
DDA – Discontinuous
 Not able to model large rotations.
solution within ‘blocky’ ground conditions.
Deformational Analysis
procedure
Coding considers crack tip stress
PFC – Particle Flow Static analysis provided, does not
concentrations
Code always provide the sophistication
that is required when considering
the complex factors controlling
initiation and eventual sliding (Stead
et al., 2006a).
42

42
CHAPTER 2: Literature Review

(Continuation of Table 2.2 - Basis of some of the more frequently used methods of analysis that are available, with suitability, benefits and limitations)

Basis of model and


Composition Suitability Benefits Limitations
method of analysis

FEM/DEM incorporates formulations to


represent the inherent Discontinuous,
Inhomogeneous, Anisotropic and Non-Elastic
(DIANE) nature of a rock mass (Elmo, 2006).
 Complex software, requiring
Complex problems that FEM/DEM can accommodate the transition from
special training
require the consideration of a single finite element domain into a number of
 Requires very specialised input
more advanced fracture interacting domains, each of which is
mechanics. represented by its own separate finite element
parameters, in addition joint
Finite/discrete mesh (Munjiza et al., 2004).
properties are required that are
elements
 BEM is most commonly
not routinely measured
All advantages of the discontinuum and
Hybrid analyses
 Requires ongoing calibration and
Continua to used for simulating far-field continuum techniques upon which they are
discontinua rocks as an equivalent based, plus FEM/DEM can:
 Can model large rotations
BEM/FEM testing of constraints (Stead et
algorithms elastic continuum,
 FEM and DEM are  Allows for new fracture propagation and
al., 2006a)
DEM/BEM
 Sophisticated models can
Can use either appropriate for the non- existing fracture extension through
explicit or implicit linear or fractured near- mesh;
 Actively simulates comminution of failed
DEM/FEM require extensive runtimes, using
solution field, where plasticity (non-
a large amount of computer
procedure as linear mechanical material
 Also dynamic, 2-D and 3-D analyses
processing power and memory
desired behaviour) or the explicit
 Groundwater module still in
representation of fractures possible with a wide variety of
is required (Jing, 2003) constitutive models (plastic, visco-
development stage.
plastic, etc.) (Stead et al., 2006a).

‘Compelling’ advantages to using the combined


FEM/DEM method, especially for modelling
problems involving large geometric changes
and post-fracture particle flow (Owen et al.,
2005).
43

43
CHAPTER 2: Literature Review

Domain, continuum and discontinuum solution procedures


A domain solution is derived from computation of the modified surrounding stresses,
via a series of calculations concerning the strength of each element and its interaction
with the strain imposed from an excavation. The resulting modified in-situ stress state
is by its nature is heterogeneous, derived ultimately from the geologic structure and
loading path which should be simulated by the numerical model (Hart, 1991).

Within continuum approaches the fundamental basis of a model can be either finite
element or finite difference, each being a distinct method of calculus. The Finite
Element Method (FEM) is one of the most popular methods of numerical analysis;
Rocscience (2004) state that FEM was first applied to geotechnical engineering in 1966
and has become increasingly popular due to its wide ability to analyse a variety of
problems. However Owen et al. (2005) describes the lack of robustness of the overall
finite element framework when the incompressibility limit of a material is approached,
the conventional displacement-based finite elements with ‘low order shape functions’
can cause a locking behaviour.

The coding within the FEM utilises a numerical approximation of the connectivity of
elements, continuity of displacements and the stresses between elements to arrive at
solution, whilst the Finite Difference Method (FDM) instead works on stress-
displacement relations and stress-strain equations to generate appropriate mass
behaviour (Eberhardt, 2003). The FDM is a technique that is very similar to FEM’s,
only it is applied to problems with simple boundaries along which a regular mesh is
constructed; subsequently short comings exist when modelling fractures, complex
boundary conditions and material inhomogeneity (Elmo, 2006). Also as a result of the
basis upon continuum mechanics, the FEM cannot simulate situations where there is
large-scale opening, sliding and complete detachment of elements (Jing and Hudson,
2002).

Even FEM codes with explicit solutions are inadequate for modelling highly non-linear
behaviour which occurs within ‘blocky’ ground conditions, instead with such situations
the discrete element method (DEM) is more appropriate (Hoek et al., 1990). The DEM
and Discontinuous Deformation Analysis (DDA) are both forms of discontinuum
modelling. The DDA was developed to perform complete deformational analysis of a
block system by Shi in 1988 (Jing, 2003). The fundamental difference of the DDA
approach to the DEM approach is that unknowns in the equilibrium equations are
replaced with displacements as opposed to forces; consequently a matrix, which is
used in a similar manner in FEM’s analysis, can be used to solve equations within DDA
methods (Eberhardt, 2003). DDA methods are well suited to the analysis of large

44
CHAPTER 2: Literature Review

deformations and rigid body movements; its ability to simulate coupling or failure states
between contacted blocks is also advantageous (Eberhardt, 2003). Block rotation,
fracture opening and complete detachments are all rigid body motions of individual
blocks within a model, the displacements caused by these motions are impossible to
model in FEM, FDM and BEM but such motions are possible with DEM techniques
(Jing, 2003).

The following sections introduce various numerical methods, some of which are
commercially available for use as a tool in slope stability analyses.

2.3.2 Kinematic Analysis – Siromodel


There are many models under development for considering kinematic failure of blocky
rock masses, for instance Moffitt and Rogers (2007) present a kinematic analysis that
is directly linked to the discrete fracture network (DFN) generated by FracMan (Golder,
2008). However their model only considered failure within an underground situation.
Siromodel (CSIRO, 2008a) is program that is being developed for kinematic analysis of
rock slopes; CSIRO (2008a) briefly describes that from input structural distribution
data, the Siromodel splits the rock mass into polyhedral blocks. Consequently Read
and Ogden (2006) describe Siromodel as a 3D computer model which has the ability
to:

 Account for the spacing, orientation and continuity data for discontinuities;
 Predict how candidate failure surfaces can propagate through the rock mass,
identifying pathways of least shear and/or tension resistance thus locating
critical rock bridges.
 Identify potential direction of failures, simulating kinematically free blocks on
curved or undulating surfaces.

Siromodel uses a DFN to create the 3D polyhedra; the DFN can be generated either
via manual or digital processes and can consequently take the data generated by
Sirovision (Read, 2008). As described in Section 2.3.1, Poropat and Elmouttie (2006)
introduce a kinematic analysis using Sirovision data. Grenon (2007) presents a similar
technique, although the 3D ‘susceptibility model’ that they produce is derived ultimately
from structural mapping data. Importantly the method used by Grenon (2007) is an
approach that integrates geographical information system (GIS) software, with
kinematic analyses, via the following process:

45
CHAPTER 2: Literature Review

1. Firstly each of the designated structural domains within the pit is split into cells
by the GIS software, which considers the slope orientation within each of the
cells, in conjunction with the discontinuity network from structural mapping.
2. Subsequently the GIS software performs a kinematic analysis within each cell,
assessing the kinematic feasibility of planar and wedge failure, to give a hazard
map; importantly Grenon (2007) suggests that statistical fracture properties are
assigned during this analysis.
3. Finally a limit equilibrium analysis is used to derive the FOS, and also the
probability of failure (from Monte-Carlo simulations), of the respective planar
and wedge failures that were identified from the kinematic method.

Siromodel is perhaps a step beyond the kinematic analyses within the process
described above. With the ability to consider rock-bridges and curved or undulating
surfaces, there is the potential of Siromodel being integrated with more advanced
forms of numerical modelling, which can consider the rock strength. However CSIRO
(2008a) outlined that Siromodel is limited to bench and inter-ramp scale, with the larger
scale rock slope strength addressed by the SRM method, discussed in Section 2.5.2.

2.3.3 Limit Equilibrium Analysis – SLIDE and other programs used


during this research
Limit equilibrium methods allow weakness within a slope to be considered through the
application of an anisotropic and non-homogeneous strength function (Moffitt et al.,
2007). This strength function can be either discontinuity or mass-related; importantly
all limit equilibrium methods involve the solution of driving and resisting forces, either
by using a principal composition of slices, or considering a discrete block.

SLIDE is a program that has been developed by Rocscience (2008), and is widely used
in the geomechanics community to assess mass-controlled slope instability. It has a
principal composition of vertical slices, with an automated search that allows the
detection of an undefined failure surface through a homogeneous material in 2D. Pore-
pressures can be inputted to allow the calculation of heads and discharges. Finally
there is also a probabilistic function within SLIDE, making sensitivity and back analysis
studies quicker.

Both circular and non-circular failures can be considered in SLIDE, although


importantly no direct discontinuity-related strength can be assigned when performing a
non-circular analysis, therefore this principally simulates a mass-controlled failure

46
CHAPTER 2: Literature Review

mechanism. There are ten, slice equilibrium analysis methods that are available within
SLIDE; for more detail on these see Appendix C.

As well as SLIDE, three other limit equilibrium methods have been used during the
research reported within this thesis. Firstly, similar to SLIDE, there is the Microsoft©
Excel-based spreadsheet Jacob (Pine, 2006a). Unlike SLIDE, at the base of each
individual slice, Mohr-Coulomb strength is assigned when using Jacob. This allows the
consideration of both mass and discontinuity strength upon a defined path. Chapter 7
introduces a failure where Jacob had to be used as opposed to SLIDE, due to this
unique feature.

Briefly two other limit equilibrium techniques have also been used, which both consider
discontinuity-related 2D failures, with the simulation of a discrete plane failure block as
opposed to a collection of slices. Plane_failure (Pine, 2006b) is a Microsoft© Excel-
based spreadsheet; this again allows a high degree of control upon input parameters.
However its equivalent, RocPlane (Rocscience, 2008), permits an angle of dip to be
specified for the tension crack if desired; this can be important when considering
active-passive type geometries as in Chapter 6. Both Plane_failure and RocPlane can
use Monte-Carlo simulation to examine the effects of parameter variability.

The following section introduces Phase2 which has been used throughout this thesis,
as a form of continuum analysis to further the understanding, which has been gained
from limit equilibrium modelling.

2.3.4 Hybrid elasto-plastic finite element (boundary) Method –


Phase2
Phase2 is a two-dimensional elasto-plastic numerical modelling program developed and
marketed by Rocscience (2008); its basis is upon a hybrid finite/boundary element
method (FEM/BEM). Rocscience (2008) describe the benefits of using both SLIDE and
Phase2 together, with the prospect of importing models from SLIDE for computation in
Phase2. This allows the direct comparison of limit equilibrium and finite element
analyses.

The hybrid FEM/BEM programming permits Phase2 to model multiple materials and
simulate non-linear behaviour close to the excavation boundaries, whereas further
a-field in-situ stresses are modelled via a boundary method (Hoek et al., 1998). This is
different to a solely BEM approach, which as Jing and Hudson (2002) suggests, is
more suited to homogeneous and linearly elastic bodies. This is because the BEM

47
CHAPTER 2: Literature Review

divides only the boundary of an excavation into elements, whereas the rock mass
within the centre is represented mathematically as an infinite continuum (Meyer, 2002).
Within this central continuum, non-linearity and inhomogeneous features of the rock
mass can not be modelled; however as Phase2 employs a hybrid approach, such
aspects can be considered, as it is only the boundaries of the model that are simulated
as a continuum.

This FEM composition of a Phase2 model allows it to simulate progressive failure/post-


failure behaviour of the rock mass and rock-support interaction (Lightfoot and
Maccelari, 1999). Phase2 also has the ability to model groundwater seepage
(Rocscience, 2008).

Importantly the consideration of discontinuity-controlled failures is limited due to the


FEM (continuum-based) approach within Phase2. Discontinuities can be specified
within the FEM mesh, with a range of terminations; however direct displacement
cannot be simulated upon these. Starfield and Cundall (1990) suggest that FEM has
difficulty simulating large movements and rotations as a result of pre-specifications of
the interconnected special elements which represent joints. Instead the qualities of
Phase2 make it more appropriate for mass-controlled shear failure, which can be
analysed using the shear strength reduction (SSR) method, a relatively new feature of
Phase2 6.0.

Shear Strength Reduction (SSR) technique:


The SSR technique permits the derivation of a critical strength reduction factor (SRF),
by which the strength parameters must be downgraded by in order to permit failure
upon an undefined failure surface. The SRF is therefore equivalent to a FOS, (see
Section 2.2.2), in that it represents the amount by which the mass strength must be
reduced, to reach the very limit of failure and therefore a FOS of 1.

During the SSR finite element technique, slope materials are assumed to have an
elasto-plastic strength behaviour; progressive reduction of the material shear strengths
then occurs until there is a solution of the subsequent collapse (Rocscience, 2004).
The resulting SRF (also referred to as F), represents a value by which the mass shear
strength of the material must be reduced by to result in failure; for Mohr Coulomb
material this equates in the following formula (note that the SSR approach is
conventionally applied to Mohr Coulomb material as a result of the ease with which the
shear strength parameters may be downgraded):

48
CHAPTER 2: Literature Review

mass cmass tanmass


= + [2.4]
F F F
Where , c and  are the shear, cohesive and frictional mass strengths; F represents
the SRF by which the Mohr-Coulomb strength components are reduced by to obtain a
reduced shear strength.

Rocscience (2004) identify several benefits of the SSR method that can be performed
in FEM techniques such as Phase2, over the conventional calculation of the factor of
safety (FOS) from limit equilibrium techniques. These are outlined below:

 The SSR method eliminates the need for the calculation of the FOS via a model
that is specifically based on slope failure via a particular mode (wedge, planar,
toppling or circular failure).
 No assumptions are required on the location and shape of the failure surface.
 Arbitrary assumptions regarding the inclinations and locations of inter-slice
forces are not required with the use of the SSR analysis.
 The SSR approach can also be more readily applied to 3D slope simulations
than limit-equilibrium models.

The fact that no a priori assumptions on the failure mechanisms are required, when
using the SSR to back analyse a slope failure, is a distinct advantage over some
kinematic and even limit equilibrium approaches which require the characterisation of
the failure mechanism prior to modelling. Hammah et al. (2007) suggest that the SSR
technique is a particular advantage for blocky rock masses where failure mechanisms
may constitute a combination of shearing through intact rock and failure along discrete
discontinuities.

However, it can be suggested that the SSR technique is only applicable for failure
mechanisms where shear failure is occurring. In situations where there is internal
separation Diederichs et al. (2007) notes that the SSR technique is less effective.
Phase2 does have the option to include the tensile strength during a SSR simulation,
reducing this by the SRF in the same way shown in Equation 2.4. Diederichs et al.
(2007) suggests that this factoring of the tensile strength is only necessary in situations
where tension is likely to be inherent in the failure process, and not in situations where
there is solely shear/sliding during failure.

Another issue with the SSR technique within Phase2, is the fact that the SRF that both
c and tan are downgraded by, has to be same. This was an issue when performing a
back-analysis of rock mass strength in Chapter 7, as it was not possible to fix one of
these and independently analyse the influence of each; this limited the comparisons
49
CHAPTER 2: Literature Review

that could be drawn between the other methods that were used. Diederichs et al.
(2007) also state that this independent application of SRF could be important, as
cohesion is often open to more variability than frictional strength; also the control of
cohesion dominates over friction for shallow problems.

Importantly, the SSR approach is typically limited to materials that follow the Mohr
Coulomb linear criteria; most of the alternative material criteria to the Mohr-Coulomb
are nonlinear, such as the Generalised Hoek-Brown and Power Curve strength models.
With the non-linear criterion it is not as easy to calculate the reduced strength
parameters during the SSR analysis, in fact Rocscience (2004) state that it is
impossible to generate such close-formed relations as those that can be created when
using linear failure criteria. As a result it could be considered that the SSR technique is
limited to application within continuum numerical models, where elasto-plastic strength
is assumed for slope materials (Hadjigeorgiou et al., 2006).

Diederichs et al. (2007) suggest that progressive failure can be captured using the SSR
technique. This is due to the ability of the technique to show progressive reduction in
strength of the rock mass. Hammah et al. (2007) state that there is a specific time
during which a SSR analysis is most effective, which is just before failure when there
are typically small displacements.

The ability of SSR approach in modelling progressive failure is clear when considering
the method in an excavation application; reduction of shear strength once the final
excavation block is removed can allow shear strain to develop, marking the location of
either a continuous potential failure plane or the linkage of step-paths. However the
consideration of step-path failure and localised rock bridge strength is inhibited in FEM
models, which is why one must look to more advanced numerical models such as
UDEC, PFC or ELFEN to model such behaviour.

2.3.5 Finite Difference (domain) Method – FLAC


FLAC (Fast Lagrangian Analysis of Continua) is a continuum numerical modelling
approach which was initially written by Peter Cundall and has been available
commercially since 1986; now it is marketed and has been extensively developed,
even into a 3D model, by Itasca (2008) and HCItasca (2008). FLAC is a finite
difference method (FDM) technique using an explicit formulation. User control is
provided through the built in programming language, FISH. Although FLAC does not
offer the potential to model large numbers of fractures, a few discontinuities with simple
intersections can be included (Lightfoot and Maccelari, 1999).

50
CHAPTER 2: Literature Review

HCItasca (2008) report on three different FLAC programs: FLAC, FLAC/Slope and
FLAC3D; all of these have the ability to include the influence of a phreatic surface.
FLAC/Slope is a development of FLAC which provides user-friendliness to slope
models that can otherwise be developed in FLAC. The numerical coding and graphical
interface of FLAC is used to deliver an ‘advanced factor-of-safety determination’ for
rock and soil slopes in 2D HCItasca (2008). FLAC and FLAC/Slope can only be
3D
applied to 2D; FLAC is a further development of FLAC, which allows the modelling in
three dimensions.

The SSR technique of strength determination, discussed in the previous section, is also
possible within FLAC. Makusha and Minney (2006) report that a SSR technique is
available within FLAC/Slope, this technique poses the benefits and disadvantages
discussed in Section 2.3.4. The following section, reviews some examples where
FLAC has been applied to fractured rock slopes; note that in order to represent a
fractured rock slope, an equivalent material has to be input into FLAC. Discussion is
given later in Section 2.5, on methods by which equivalent materials can be created in
order to represent the influence of discrete features.

Some applications of FLAC to rock slope problems:


There have been several International Symposia based on FLAC and its application to
numerical modelling in geomechanics. During the Second International FLAC
Symposium Poisel et al. (2001) presented a FLAC3D model, to calculate displacements
on the main through-going joints on a rock slope in Austria.

The Third International FLAC Symposium in 2003 saw applications of FLAC to circular
type failures, with the calculation of FOS using the shear strength reduction technique
within FLAC and comparisons drawn against the FOS calculated by limit equilibrium
methods. During the Fourth International FLAC Symposium Preh and Zapletal (2006)
presented a study on the influence of mesh size on the solutions predicted by FLAC;
demonstrating that the FOS decreased with increasing mesh resolution thus with a
coarser mesh the stability of a slope could be overestimated.

Handley and Karparov (2007) modelled an active/passive type mechanism within an


open pit coal mine. In their case, FLAC did not predict the appropriate shear
behaviour; instead only tensile failure was well represented.

Makusha and Minney (2006) used FLAC/Slope to study the influence of changing
batter angles, soil thickness and spoil loading on the stability of coal strip mine slopes;
however they noted limitations in the ability of FLAC/Slope to represent multiple
discontinuities and consequently used UDEC for fault and dip angle sensitivity studies.
51
CHAPTER 2: Literature Review

Board et al. (1996) produced a FLAC model of Chuquicamata indicating a global


slumping mechanism due to toppling from high angle joints. They compared the
results from FLAC and UDEC models of Chuquicamata; more displacement at the
benches was predicted in the UDEC model where shear on joints is kinematically
possible. This is due to the particular qualities of UDEC, which are discussed within
the following section.

2.3.6 Distinct Element (domain) Method – UDEC and 3DEC


Universal Distinct Element Code (UDEC) and 3DEC are both DEM approaches
developed by Itasca (2008), which use an explicit solution scheme, with UDEC able to
model in 2D and 3DEC in 3D. User control is provided through the built in
programming language, FISH. UDEC can be considered to fall within the second level
of sophistication outlined by Stead et al. (2006a), see Section 2.3.1. This is because
the code does not have any re-meshing capability; consequently UDEC is unable to
directly simulate kinematic release; however brittle fracture propagation can now be
modelled using the Voronoi Tessellation scheme discussed later. Without Voronoi
Tessellations, UDEC can only predict movement and deformation upon through-going
discontinuities that are inserted into the model. Although, the fact that UDEC is able to
represent the discontinuities explicitly, allows the study of relative shear and opening of
joints (Board et al., 1996).

Numerical models which include discrete fractures can consist of deterministic and
stochastic components. The stochastic component constitutes the portion of the rock
mass that has not been directly sampled but instead comprises fractures that have
been generated statistically, to have the same statistical distributions and geological
correlations as the deterministic explicitly mapped discontinuities (Kleinei et al., 1997).
Statistical modelling has been conducted at Camborne School of Mines, through the
research of both Elmo (2006) and Ford (2008); the former provided a case study on
which rock mass strength determination was based, in Chapter 8.

DEM models comprise an assemblage of deformable blocks separated by joints of


known stiffness (Eberhardt, 2003); as a discontinuum method the DEM reflects the
component movement of a system (Elmo, 2006). Consequently DEM’s are particularly
powerful for blocky rock masses, modelling the rock mass as an assemblage of rigid or
deformable blocks; therefore they can predict large displacements, rotations and
complex constitutive behaviour for intact material (Kulatilake et al., 1995).

52
CHAPTER 2: Literature Review

As outlined in Table 2.2, an analysis based upon a DEM is complex, with experience
required to understand relevant modelling procedures and input parameters. Due to
the complex inputs and sensitive environment of a DEM solution, discontinuum
analysis can be very sensitive to change. Starfield and Cundall (1990) suggest that
different results can arise from a change in the initial conditions, or even if other factors
are changed, such as the type of computer! Hart (1991) also relates to the sensitive
nature of a DEM model, implying that a typical problem in DEM models is the locking
up effect, which is associated with incorrect contact properties.

Moffit et al. (2007) describe a useful feature of UDEC that allows the consideration of
discontinuity statistics, with an option to enter mean values for fracture set dip,
orientation and the gap length between consecutive fractures of the same length.
However they found that this approach was limited to a relatively large scale study as
numerical instabilities were discovered when using small block sizes due to irregular
block geometry.

Another relatively new addition to UDEC, is the ability to consider fracture via the
Voronoi Tessellation joint generation scheme. Alzo'ubi et al. (2007) report on a UDEC
damage model (UDEC-DM), describing the process to form a mesh of randomly sized
polygonal blocks. A selection of these polygonal segments can be assigned properties
to dictate the strength of the bonds (fractures) between the blocks. Subsequently
tensile or shear fracture through intact rock within designated areas can be studied,
with UDEC simulating the formation of discrete fractures. However from only
preliminary assessments, Karami et al. (2008) suggest that the analysis of rock-bridges
in UDEC via the Voronoi method is very computationally demanding.

Some applications of UDEC and 3DEC to slope problems:


There are many examples of where UDEC has been applied successfully to rock slope
problems; 3DEC has had fewer applications. An example of a back-analysis is the
UDEC research by Coggan and Pine (1996) applied to the Delabole failure that
occurred in 1967. In this case UDEC aided the understanding of the progressive
failure mechanisms involved, identifying the driving mechanism to be a “chisel effect”
as a result of the interaction between four major blocks. Coggan and Pine (1996) did
find limitations in their distinct-element modelling of the Delabole failure, with UDEC
being unable to effectively consider the progressive-time dependent weakening that
occurred during the progressive failure within the slope. In addition, the influence of
rock-bridges, which exist between the discontinuities, could not be appreciated during
their UDEC simulations.

53
CHAPTER 2: Literature Review

Mitani et al. (2004) used UDEC to study flexural toppling; their findings proved there is
a critical joint dip angle and that joint spacing influences the likelihood of toppling. For
instance if the joint spacing is reasonably large then the inclination of the slope cutting
can be larger before flexure toppling occurs.

Alejano (2004) used UDEC to illustrate the theoretical failure within footwall slopes of
an open pit mine where there are reasonably closely spaced discontinuities parallel to
the cut of the slope. This allowed the development of a limit equilibrium approach and
the creation of a calculus method to determine the FOS of each of the failure
mechanisms: bilinear slab failure (Figure 2.6a), and ploughing slab failures (sliding and
toppling modes, Figures 2.6b and 2.6c respectively). These were compared against
analyses within UDEC.

(a) (b) (c)

Figure 2.6: Example of fully joint controlled failure mechanisms analysed in UDEC (after Alejano,
2004).

UDEC can be used to investigate step-path problems. Moffitt et al. (2007) presents a
UDEC SSR approach within a problem that requires consideration of step-path failure.
A major limitation of such an approach is that the failure surface is highly dependent on
block size, as when using UDEC fractures have to terminate on other fractures,
consequently shear through the rock mass cannot be simulated (unless using the
Voronoi Tessellation scheme), and therefore overly conservative factors of safety could
be derived.

As stated previously, 3DEC has had fewer applications to rock slopes, most probably
due its complex nature. Reference was given in Section 2.2.5, to Yan et al. (2007b)
who used 3DEC to model step-path failure. Sainsbury et al. (2007) and Brummer et al.
(2006) both give examples, where large slopes or whole open pits have been modelled
using 3DEC. The derivation of mass strength properties, for programs such as 3DEC,
is discussed in Section 2.5; one of the ways is to use Particle Flow Code (PFC) to
derive an equivalent mass strength. PFC is introduced within the following section.

54
CHAPTER 2: Literature Review

2.3.7 Particle Flow Code


PFC is a DEM consisting of many circular (2D) or spherical (3D) particles with the
ability of particles to cluster where there are joint bounded blocks, and break apart via
stress-induced breaking of bonds between the particles where discontinuities extend
through the medium; the interaction of the particles is through frictional sliding contacts
(Eberhardt, 2003, Stead et al., 2006a). PFC was originally written by Peter Cundall
and released it as a commercial code in 1995 for application to block caving (Itasca,
2008); now 2D and 3D formulations are available.

Eberhardt (2003) outlines how PFC can be applied to study joint bounded blocks on a
macro scale, through to grain-to-grain contacts on a micro level. Therefore internal
slope deformation, due to yielding and intact rock fracture of jointed rock masses, can
be studied using PFC. Stead et al. (2006a) suggested that this was a significant
development upon the previously less comprehensive distinct element methods.

Intact rock breakage associated with slope failures, can be modelled by PFC codes via
the breakage of assumed bonds between particles or spheres (Stead et al., 2006b).
Read (2007) describes how PFC simulates fracture by the breaking of individual bonds
between particles and coalescence of these to form microcracks, although importantly
this is not based upon a macro-mechanics fracture criterion such as Mohr-Coulomb,
instead PFC uses micro-mechanics without basis on a particular criterion.

Mas Ivars et al. (2007) state that PFC3D now possesses a (‘parallel-bonded’) stress
corrosion model, which is able to simulate time-dependent behaviour. They also
suggest that other elements of time-dependency, are still to be considered, such as
joint-slip, crack healing, cementation and densification. Also, a pore pressure coupling
is still to be developed.

Within PFC the rock mass is composed of an amalgamation of many small spheres as
opposed to distinct irregular shaped blocks, which ELFEN can form (introduced in
Section 2.3.8). Fakhimi (2004) discusses some of the limitations of PFC, identifying
fundamental issues regarding the accuracy of the frictional angle, and the ratio
between unconfined compressive strength and tensile strength, when materials are
modelled using circular particles. Another issue regarding the composition of particles,
is the representation of a discrete joint. Lorig (2007) referred to the ‘smooth contact
model’ that has been introduced into the PFC code, as a result of the otherwise very
irregular unrealistic joints that can be generated in PFC. A brief discussion is given
below concerning the application of PFC to fractured rock slopes.

55
CHAPTER 2: Literature Review

Some applications of PFC to rock slope problems:


There are relatively few published examples of rock slope analyses using PFC. The 1st
International PFC Symposium in 2002 saw the application of PFC to rock fall analysis
and large-scale landslide modelling. Also a failure mechanism between a hard
competent rock, upon a soft incompetent base, were investigated using PFC2D; this
demonstrated the sliding/slumping failure of a large landslide that acts almost as a
continuum.

More recent applications of PFC, has been through the Large Open Pit (LOP) project
(further discussion on modelling approaches within the LOP, is given in Section 2.5.2).
Read (2007) and Lorig (2007) report on a PFC model of a large slope, they both noted
limited fracturing across rock-bridges; this was attributed to being an artefact of the 2D
analysis, as discussed in Section 2.3.9. The following section reviews ELFEN, the
numerical package that has been used during the progress of this research.

2.3.8 Hybrid finite/discrete element (domain) Method – ELFEN


The application of a hybrid FEM/DEM numerical method combines advantages from
both discontinuum techniques and continuum approaches. This allows the modelling
of intact behaviour, interaction along existing discontinuities, and the initiation and
development of new fractures via consideration of key principles of fracture mechanics
(Stead et al., 2006a). In addition, past applications of ELFEN to rock slopes has
demonstrated its ability to model kinematic release, complex internal distortion and
dilation.

Primarily within ELFEN the rock slope is represented as a continuum using finite
elements. In the process of progressive fracturing, ELFEN then forms fracture-
bounded blocks, which are deformable finite elements; consequently ELFEN has the
ability to model a slope failure from initiation, transport and comminution to deposition
(Stead et al., 2004).

The adaptive remeshing scheme and contact search algorithms in ELFEN, allows the
simulation of brittle fracture initiation and propagation through the finite element mesh
(Eberhardt, 2003; Stead et al., 2006a). As a result of this remeshing capability, hybrid
methods are often applied to situations that are highly dynamic with rapidly changing
domain configurations (Owen et al., 2005). Within ELFEN the path of fracturing can
pass through elements (intra-element fracturing), or along element boundaries (inter-
element fracturing). The intra-element fracturing capability can pose as a significant

56
CHAPTER 2: Literature Review

advantage, as no pre-designation of fracture path is required. However it has come to


light during the progress of this research, that intra-element fracturing can cause
numerical instability and consequent excessive runtimes due to the small ‘badly’
shaped elements, formed during present re-meshing scheme. This considerably limits
the beneficial fracturing mode within ELFEN.

Perhaps one of the most important and unique aspects of ELFEN is its ability to
seamlessly simulate the transition of a rock from continuum to discontinuum states (Cai
and Kaiser, 2004). Effectively the FEM component of ELFEN is utilised until the yields
occurs. Once yielded, the stress state within the material subsequently succeeds the
fracture criterion; as a result a crack is initiated through the DEM aspect of ELFEN.

The dual meshing ability of ELFEN allows the intact joint bounded blocks to be
represented by a finite-element mesh, whereas the joint behaviour is modelled using
discrete elements. This allows the behaviour within a rock slope prior to and during
failure to be studied, with the simulation of fracturing, damage and associated softening
(Stead et al., 2006a).

Finally, because hybrid codes are the most advanced level of numerical analysis, it
consequently requires the highest intensity of field mapping. Stead et al. (2006a)
consider the ‘total rock slope failure,’ suggesting that it is necessary to collect data from
all areas of the failure; this includes zones of initiation, transportation and deposition.
This level of detail would constrain the degree of comminution which is predicted by
ELFEN, providing further verification of the frictional flow of discrete elements that
ELFEN presents during a slope analysis.

Like most other comprehensive numerical models that have been discussed within this
section, 3D simulations can be performed within ELFEN. To take further opportunity of
this feature, an advance in computer power is needed to allow the level of discrete
detail to be simulated. The compromise of 2D simulations of fractured slopes is
discussed in the following section.

2.3.9 Comparison of two-dimensional and three-dimensional


analyses
3D models provide a more realistic representation of actual slope problems than 2D
simplifications. However as Hart (1991) suggests, a 3D distinct element model
requires vast calculations; consequently their use can be limited by the memory
capacity and calculation time restrictions of modern-day personal computers. Prior to

57
CHAPTER 2: Literature Review

2003, the 3D analysis of slope problems was an uncommon practice (Wyllie and Mah,
2004). However developments on the 3D facilities provided within programs such as
FLAC, 3DEC, PFC and now ELFEN make 3D numerical analyses a more accessible
tool.

Eberhardt (2003) consider 3D models to be particularly suited to the analysis of wedge


instabilities and also the examination of the influence of rock support instalments.
Wyllie and Mah (2004) suggest that there are certain situations where a 3D analysis is
recommended, this includes slope problems where:

 The principal geological structures or the axis of material isotropy does not
strike within 20-30° of the slope;
 Principal stresses within the slope are neither perpendicular or parallel to slope
strike;
 And where there is a variation in geology and consequent geomechanical
behaviour, along the slope strike.

Another situation which would benefit from a 3D analysis is step-path failure. As


reported in Section 2.3.7, a large slope PFC model revealed limited rock-bridge failure.
Both Read (2007) and Lorig (2007) attribute this to a limitation of their 2D analysis; with
a lower likelihood of closed volumes than would potentially form within a 3D analysis.

As presented in Section 2.2.2, Wyllie and Mah (2004) suggest that the curvature can
have a significant positive effect on stability in massive rock/non-structurally controlled
slopes; this is an important aspect that is not addressed directly via 2D analyses.
Instead a 2D analysis comprises of a unit slice through an infinitely long slope with the
assumption that the radii of both the slope and toe are infinite (Wyllie and Mah, 2004).

As a result of their greater authenticity ideally 3D analysis should be carried out in


preference to a 2D analysis, where possible. It has been found that there are certain
slope situations where a 3D analysis is recommended, although at present detailed
analysis are still inhibited by computing speed. Despite certain limitations of 2D
analyses, they can potentially act as an adequate and realistic simulation of real slope
problems, presenting a simplified yet representative model.

58
CHAPTER 2: Literature Review

2.4 Groundwater within rock slopes


One of the most important controls on slope stability is the groundwater conditions
within the slope; reduced normal stresses within discontinuities and reduced shear
strength can be the result of poor drainage within a slope.

Within this particular section, there are two parts:

1. Firstly there is a review on how groundwater affects rock strength. The primary
discussion within this part gives an appraisal of groundwater processes at the
larger slope-scale. Subsequently groundwater flow within rock slopes is briefly
reviewed, with a final discussion on the smaller scale, reviewing the influence of
groundwater on the brittle fracturing of rock.
2. Finally the particular ability of the newly developed groundwater module, within
ELFEN, is outlined.

The subsequent aim is to develop an understanding of the key aspects of groundwater


presence that must be considered when constructing a numerical model.

2.4.1 Influence on overall slope stability


The influence of pressure from groundwater on rock strength is substantial;
consequently the accurate understanding of groundwater distribution within a rock
mass, and the correct simulation of the principal effects of groundwater within
numerical models, is very important. In the presentation given by Johnson et al. (2007)
it was suggested that groundwater can have the greatest influence on the design of a
slope. For instance the Equation 2.1 (given in Section 2.2.2), to calculate the
inclination of the critical failure plane, is not justified for situations where groundwater is
present. Instead the inclination of failure plane through intact rock, as calculated
through Equation 2.1, can be up to 10% lower if water is present within a tension crack
(Hoek and Bray, 1981).

Groundwater can influence rock strength in a number of ways, mainly though the
additional stress from groundwater promoting brittle fracturing. At the small scale the
saturation state of an intact rock has significant influence on its strength, most probably
due to additional pressure within microcracks. Romana and Vasarhelyi (2007) suggest
some tentative rules of thumb for preliminary reduction of strengths when dealing with
saturated samples (although it must be noted that the relationship between saturation
and strength is not linear):

59
CHAPTER 2: Literature Review

 Strong rock (indurated): UCSdry/UCSsat = 0.8 to 0.9


 Medium strength (cemented) rock: UCSdry/UCSsat = 0.6 to 0.7
 Soft (argillaceous) rocks: UCSdry/UCSsat = 0.3.

At a larger scale, more mechanical principles have to be considered to fully appreciate


the influence of groundwater. Not only does the enhancement of brittle failure have to
be considered within a numerical model, but also the flow through the fracture network
and increased fluidity of moving particles controlling post deformational behaviour.
Consequently coupling of rock mechanics with other disciplines is necessary. Sullivan
(2007) lists four types of hydromechanical coupling that demonstrate the influences
groundwater can have on a rock mass:

1. Direct solid to fluid coupling – changes in pore fluid pressures are generated by
a change in stress;
2. Direct fluid to solid coupling – where a change in rock mass volume is caused
by a change in pore fluid pressure;
3. Indirect solid to fluid coupling – an alteration in hydraulic properties of rock
mass is a consequence of a change in stress; and finally,
4. Indirect fluid to solid coupling – by which mechanical properties of the rock
mass are modified as a result of pore fluid pressure.

An illustration of how mechanical principles interlink during the progression of a slope


failure, is given in Figure 2.7.

An open pit mine can have unique hydrological characteristics; the reader is referred to
Sullivan (2007) for more detail. Importantly there are two hydrological areas where
water can be present within an open pit slope that do not occur within a natural slope:

1. A “fractured rock aquifer” can result parallel to the pit walls due to disturbance
from blasting,
2. Also unloading of the rock mass during the excavation of the slope can
increase hydraulic conductivity and storage within a zone surrounding the open
pit (Sullivan, 2007).

60
CHAPTER 2: Literature Review

Figure 2.7: Illustration of different areas of mechanics that must be considered when
modelling a slope failure from initiation through to deposition (after Sullivan, 2007).

Groundwater may affect slope stability on a number of scales; each of these has an
independent set of fundamental processes and subsequently different assumptions
have to be made for each scale of study. On the bench-scale, failure mechanisms are
often very sensitive to local rainfall, joint aperture and continuity; Sullivan (2007)
suggests that as a consequence of this, prediction of slope performance on small-scale
wet slopes is practically impossible. On a multiple bench or inter-ramp scale there are
often several structures involved in the failure mechanism; the derivation of correct
pore water distribution on each of these can also be problematic, although potential
trigger mechanisms related to groundwater are less sensitive to local variation than at
bench-scale. Finally on a large slope scale, local variation is likely to be insufficient to
influence the potential large-scale mechanism that is of interest; therefore modelling
wet slope stability situations at this scale could be considered to be more accurate than
analysing the effect of groundwater on smaller (bench) scale fractured rock slopes.

For an increased understanding of the effect that groundwater can have on slope
stability at all scales, one must consider the influence of pressure from groundwater on
shear strength and fracture mechanics in more detail. This is discussed within the
following sections.

Groundwater flow within a rock mass


The porosity and permeability of a rock mass determines the ability of the rock mass to
store and transmit water, these factors therefore need to be considered to understand
the effect that groundwater can have on slope stability. The definition of a distinct
phreatic surface, below which there is fully saturated ground, is perhaps a paradox for
a low porosity (and possibly high permeability) fractured rock mass. Instead
groundwater is held within interconnecting discontinuities, and therefore a perfect
61
CHAPTER 2: Literature Review

straight groundwater surface is somewhat of a simplification. This therefore questions


the basis of simple models, with accurate monitoring and hydrogeological modelling
necessary for a more comprehensive model.

Quadros (2007) briefly outlines the different types of groundwater flow; the most
important of these are flow within rock matrix and open fractures. Detailed research
has been conducted on fluid flow within the Carmenellis Granite in Cornwall during the
Hot-Dry Rocks Project (Pine, 1986); this provides an example of fluid flow within a
competent (crystalline) fractured rock mass. Groundwater flow within crystalline
fractured rock masses, such as the Carmenellis Granite, is a result of fractures as
opposed to matrix permeability and porosity control. Brace (1980) suggests that as a
result of fractures, in-situ permeability within a crystalline rock mass is much higher
than laboratory (small-scale) permeability, of specimens from the same lithology.

Pine (1986) lists eight key parameters required during detailed joint description for the
development of a flow model: shear strength, stiffness, dilation, effective aperture,
contact areas, roughness, tortuosity and normal compliance. With the present
computing processor power, it would be too demanding to consider this much detail
within a fractured rock mass, besides which the characterisation of these parameters
within a body of rock can be problematic. Also the flow of groundwater within a
fractured rock slope is only of interest in terms of groundwater pressure dissipation
following fracture. Such concepts have not been included within current version of
ELFEN, as discussed further in Section 2.4.2.

The propagation of fractures within an effective stress environment


The primary influence of groundwater within a discontinuity is the pore pressure ()
which reduces the normal stress (σn) acting across the discontinuity. This is given by
the ‘effective stress’ (σn'):

σn' = σn -  [2.5]

With effective stress defined, it can be included into the relationship that exists between
shear strength and normal strength:

 = c + σn' tan  [2.6]

In addition to the reduction of normal stress, groundwater can also decrease the c and
; this can be referred to as a ‘moisture effect.’ Primarily the moisture effect can
influence the intact strength of a rock. As Hoek and Bray (1981) suggest, most hard

62
CHAPTER 2: Literature Review

rocks remain unaffected by this aspect of groundwater. However the moisture effect in
softer material can be significant, in which case it is necessary to derive new c and 
properties from saturated laboratory tests.

Not only can Equation 2.6 be applied to shear strength on discontinuities, but also to
overall mass strength. In the latter case, the influence of groundwater has to be
considered in four ways:

1. The decreased normal stress (shown in Equation 2.5) on discontinuities.


2. The potential of decreased normal stress between the component grains of the
intact rock (if the rock matrix is sufficiently porous and permeable).
3. Possibility of decreased cohesive and frictional strength of the rock mass (if the
moisture effect has a significant influence on material strength).
4. Potential decreased cohesive and frictional strength of the any infill material
within discontinuities.

All of the above influences can provide a driving force that may encourage brittle
fracture. However this force may only be transient, due to groundwater flow through
the rock mass.

Stead et al. (2007) suggest that the relationship between ground water and brittle
fracture is complex with two-way functions:

1. Displacement along fractures become more probable;


2. This can cause new brittle fracture which then acts to temporarily decrease
pore water pressure by increasing permeability.

Starfield and Cundall (1988) also discuss this aspect, stating that as joints dilate during
displacement, this in turn increases the permeability; this can cause a temporary
decrease in pore pressure within the system. Subsequently as pore pressure re-
equilibrates and increases, shear strength is reduced (following Equation 2.6), causing
further fracturing.

As discussed within the following section, the current version of ELFEN is unable to
simulate the dissipation of pore pressure due to fracturing. Section 2.2.5 outlined that
progressive fracturing occurs over time; during this time pore pressure should
equilibrate. Therefore using the present ELFEN version, simulations of slope instability
can only be conducted where fracturing and consequent failure occurs as a direct

63
CHAPTER 2: Literature Review

result of saturation. The present ELFEN version could be considered inappropriate


where progressive weakening and eventual failure occurs due to fluctuations in
groundwater. This is due to the lack of pore pressure dissipation function within
ELFEN, as outlined in the following section.

2.4.2 Groundwater within a fracture-based code


The implementation of groundwater within a fracture-based code is obviously a difficult
but valuable process. At present the only other candidate software other than ELFEN,
which is increasingly being applied to fractured rock, is PFC. As outlined in Section
2.3.7, a groundwater module is yet to be developed for PFC.

Section 1.2 describes how this thesis forms part of a project, where a key aim was for
the industrial partners Rockfield Software Ltd. to develop a groundwater module for
ELFEN, which is subsequently tested during the scope of this research. The
groundwater development version of ELFEN was received just prior to the final (third)
year of this research; therefore only a limited amount of testing could be performed.

As outlined in the previous section, there are several ways in which groundwater can
influence shear strength, both within individual discontinuities and of the mass. The
current version of ELFEN can simulate the reduction of normal stress within
discontinuities. The moisture effect on intact strength is yet to be included; as is the
dissipation of pore pressure once failure occurs. Therefore the present groundwater
module within ELFEN is limited to the simulation of failure, due to pore pressure, within
only dominantly discontinuity controlled slope case studies.

With the fracture-capability of ELFEN, a simulation can be performed where an


accumulation of pore pressure within discontinuities, triggers mode I deformation and
subsequent brittle fracture extension. To test this, groundwater has been implemented
within three case study slopes during the progress of this research:

1. Within a small-scale planar failure, in Chapter 4, demonstrating the successful


decrease of discontinuity shear strength due to a rise in the phreatic surface.
2. In a simulation of a step-path failure in a chalk cliff, within Chapter 5; this has
shown ELFEN to be effective in simulating successive fracturing across of rock-
bridges due to a rise in pore pressure.
3. Finally, the simulation of an inter-bench scale failure (in Chapter 6), which has
also demonstrated that fracturing can result from a rise in the phreatic surface.

64
CHAPTER 2: Literature Review

It is suggested that although the groundwater module within ELFEN is a unique feature
amongst alternative geotechnical software, further development work is required to
improve user friendliness and create a fully coupled version. However, primarily
through Chapters 4, 5 and 6, the groundwater aspect has shown itself to be a
promising valuable asset to fracture-based geotechnical simulation.

2.5 Methods of rock mass strength determination for slopes


In this section the current methods are reviewed, which can be used to characterise
and represent rock mass strength, subsequently allowing an appraisal of slope stability.
At present there are few techniques to derive a representative strength for a rock mass;
a fundamental aspect, which was voiced during the 2007 11th Congress of the
International Society for Rock Mechanics, was the importance of rock mass properties
upon which numerical models are based. To any numerical modeller the derivation of
such properties is a difficult and questionable task, with limited time and resources that
can be available, rock mass properties are often obtained from literature (past testing
on similar lithologies), or even derived from the many empirical relationships within the
field of rock mechanics.

The derivation of rock mass properties remains one of the most difficult tasks in rock
mechanics; Cai and Kaiser (2007) suggest that field tests (plate-load, in-situ block
shear tests) form the traditional and unfortunately less practiced methods of parameter
evaluation, possibly due to cost. Instead they propose that modern techniques of back-
analysis are now more utilised, but importantly these are dependent on the constitutive
models that are adopted.

Rock mass classification schemes are perhaps the most popular approach to
characterisation. However, with the increasing scale of modern day open pit slopes, it
is becoming more common to assess situations beyond the finite empirical
envelope/knowledge-base on which classification systems have been established.
Therefore empirical-based techniques are becoming restrictive.

One of the aims of this research is to further the application of numerical modelling to
rock mass strength determination (in particular through the FracMan-ELFEN
approach). The benefit of numerical techniques over empirical-based methods is that
numerical models can simulate discontinuous aspects of the rock mass in a
quantitative manner. It must be noted that both methods can be used together, i.e.
rock mass properties can be derived from a classification-based system for use in a
65
CHAPTER 2: Literature Review

numerical model where the different scenarios of slope design can be simulated.
A distinction must be made between this type of numerical model, which incorporates
parameters based on empirical relationships, and a rock mass model that is purely
based upon intact strength data and the details of the discrete fracture network (DFN).

Creating an equivalent material has been, and still is, a common way of downgrading
mass strength in order to represent the influence of discontinuities. The two common
ways of representing an equivalent mass is either via a continuum or by including
ubiquitous joints. Board et al. (1996) describe ubiquitous-joint models as applicable to
cases where there are sets of weak continuous joints and a weak strength parallel to
the direction of the structure; consequently this allows the inclusion of directional and
weakening effects of joints without modelling them explicitly. This is often an approach
that is used within FLAC, which was introduced in Section 2.3.5. This method is limited
to situations where there are near-continuous discontinuities, as Starfield and Cundall
(1990) suggest; ubiquitous joints cannot capture block interaction and internal
movements.

Another way of representing mass strength, is to create an ‘equivalent structure;’


Schellman et al. (2006) present a stereographic way in which the joint sets are
combined to create the equivalent structure, therefore this can be performed to
average out any uneven failure surfaces. This can be considered a form of ubiquitous
joint, although it is singular and not a set of joints; however akin to ubiquitous joints, a
continuous feature is inserted that represents rock mass strength in an equivalent form.
Schellman et al. (2006) used this method of equivalent structures, to represent an
uneven bench-scale rupture surface, as illustrated in Figure 2.8.

Rock-bridge strength can be considered using the equivalent structure by simply


increasing the cohesive strength upon the through-going discontinuity; Moffitt et al.
(2007) suggest this as an approach within limit equilibrium and DEM models. A useful
example is presented by Karami et al. (2008), who identified the percentage of rock-
bridges within a DFN and proposed a relationship by which the cohesive and frictional
strength on equivalent discontinuities can be accordingly scaled (increased).

66
CHAPTER 2: Literature Review

Figure 2.8: Formation of an equivalent structure, which represents


mass strength, and consequently simulates step-path failure (after
Schellman et al., 2006).

Instead of specifying a pre-defined equivalent structure, as is depicted in Figure 2.8, an


appropriate simulation can be used to allow the failure plane to develop more naturally.
The difference is that the material, through which the mass-controlled failure occurs, is
an equivalent continuum. It is termed an equivalent continuum because it has rock
mass properties that have been appropriately downgraded, to incorporate the
geomechanical behaviour of the associated DFN. This method of representing mass
strength has been followed in Chapter 8 using the FracMan-ELFEN approach outlined
in Section 2.5.2.

The following two sections outline how classification-based methods and numerical
modelling can be used to arrive at a strength estimate. The first section gives an
introduction to the different techniques of classification-based methods, with detail on
one particular method, where an empirical-based classification system can be
incorporated within the Hoek-Brown criterion. The final part of the following section
outlines the most convenient approach of slope-design (mass strength assessment),
the Hoek and Bray chart-based method.

67
CHAPTER 2: Literature Review

Finally Section 2.5.2, provides discussion on numerical approaches of rock mass


strength derivation, through three parts:

1. Firstly a new approach for rock slope mass strength derivation is detailed, and
subsequently followed in Chapter 8, (FracMan-ELFEN technique).
2. The second part presents a similar numerical-based method for rock mass
strength, from literature where it is referred to as the ‘synthetic rock mass
(SRM)’ approach. This uses Itasca codes (PFC/UDEC).
3. Finally there is a brief outline of how the resulting equivalent continuum can be
modelled in a more simple FEM or limit equilibrium solution, to reveal slope
mass strength.

2.5.1 Classification and chart-based methods


The strength of a homogeneous rock mass can be assessed using conventional
classification based methods, which quantify geological observations and arrive at a
rating for the rock mass that is based on empirical knowledge. However the
experience, on which such techniques are based, is becoming limited considering the
scale and the complex mass behaviour that is possible in modern open pit mines. For
instance Stacey (2006) suggests that the basic technology for determining the rock
strength aspects has changed little since the 1970’s. Also during the presentation of
their paper, Franz et al. (2007) suggest that the current systems of rock mass
classification are insufficient in assessing the failure mechanisms that can occur in
large slopes.

Despite these facts, classification-based methods are more widely used than numerical
techniques, as a result of the certain advantages they have over numerical methods.
For instance, classification-approach within empirical schemes makes them relatively
simple and indirect, without an assumption of the failure mechanism. Also importantly
the geotechnical inputs, upon which empirical methods are based, can be readily
collected during mining operations.

However, one must appreciate that the classification of each input parameter can have
a significant influence on the resulting rating that describes the rock mass quality; also
the assessment of properties is qualitative and therefore can be subjective.
Consequently the use of classification approaches improves with experience, adding to
the empirical foundation on which respective schemes are based.

68
CHAPTER 2: Literature Review

For detail on the empirical rock mass classification schemes that are currently
available, see Appendix E. An important limitation, of all empirical-based systems, is
that there is no direct representation of structure within the rock mass, which can have
a significant influence on slope mass strength considering the environment of low
confining stress. Aspects from current research on step-path and multiple-plane failure
within large slopes (Elmo et al., 2007b, Franz et al., 2007), are therefore limited to
numerical approaches.

Out of all of the classification systems listed in Appendix E, the Geologial Strength
Index (GSI) and Bieniawski’s Rock Mass Rating system (RMR) system, are the most
widely used techniques as a result of their direct use in the Hoek-Brown criterion. Both
systems represent geological observations, providing a rating which downgrades the
intact strength within the Hoek-Brown criterion to present an estimation of rock mass
strength. This is described in the following section.

Classification systems coupled with Hoek-Brown Criteria


The Hoek-Brown failure envelope is a non-linear criterion, which can either be used to
describe intact strength, or equivalent continuum mass strength. It is defined by
constants σc, m and s:

 1   3   ci m  3  ci  s
' ' '
[2.7]

Where  1 and  3 are the maximum and minimum principal stresses at failure and σci
' '

is the compressive intact strength.

There are two parts to the Hoek-Brown criterion. Firstly, there are the components that
directly refer to the intact strength:

 σci (compressive intact strength)


 σ1 and σ3 which respectively refer to the maximum and minimum principal
stresses at failure

The second part to the criterion are Hoek-Brown constants (m and s), which can take
the characteristics of discontinuities within the rock mass into consideration. For intact
rock m can be referred to as mi, whereas when considering the rock mass strength, mb
is used.

As Hoek and Marinos (2006) suggest, this was neither a new nor unique approach,
instead the significant contribution was that the Hoek and Brown equation was a

69
CHAPTER 2: Literature Review

dimensionless equation that allows scaling relative to geological observations.


Subsequently the criterion was linked firstly to Bieniawski’s RMR system, with a
distinction between disturbed and undisturbed rock masses (see
Table 2.3). This tolerates the use of the criterion for both its original intent, in confined
underground settings, and also in near surface environments (which are often
disturbed), when assessing slope strength.

The RMR was linked into the basic Hoek-Brown criterion (Equation 2.7) through the
definition of the Hoek-Brown constants, as shown in Table 2.3. In addition an
association can be drawn between the RMR and the Young’s Modulus of the material,
but for only undisturbed rock masses, as shown in Equation 2.8.

E  10
 RMR 10  40 
[2.8]

Table 2.3: Derivation of the Hoek-Brown constants for disturbed and undisturbed rock masses, based on
RMR. (Equations from Hoek and Marinos, 2006).

mb mi  exp   RMR  100  14  mb mi  exp   RMR  100  28 


Parameter Disturbed rock masses Undisturbed or interlocking rock masses

s  exp   RMR  100  6  s  exp   RMR  100  9 


m

A further revision of the original criterion presented in Equation 2.7, gave the
‘Generalised Hoek-Brown criterion’ (Equation 2.9), which uses another constant, ‘a,’ in
place of the square root in Equation 2.7. This was designed to address the optimistic
rock mass strength conditions, which were otherwise predicted in poor quality rock
masses. Hoek and Marinos (2006) suggest that the introduction of ‘a’ allows the
curvature of the failure envelope to be altered, reducing the tensile strength at low
normal stresses.

 
 
 s
a

 1   3   ci mb
'

  ci 
' '
3
[2.9*]
 
*Note that this equation can also be applied to the strength of intact rock, in which case
mi is used in place of mb and an exponential of 0.5 is used in place of a.

Hoek and Brown (2002) refer to the Bieniawski’s RMR system being inadequate in very
weak rock masses. As an alternative the GSI system was developed, which performed
better in less competent rock and allowed assessment of the rock mass, (as disturbed

70
CHAPTER 2: Literature Review

or undisturbed), to be performed within the field. Equations 2.10 to 2.12 show how the
GSI was incorporated.

 GSI 100 

mb  mi exp
 
 28 14 D 
[2.10]

 GSI 100 

s  exp
 
 9 3 D 
[2.11]

a  e e 
1 1  GSI /15 20 / 3
[2.12]
2 6

As shown in Equation 2.10 and 2.11, the derivation of both s and mb requires the
definition of a further parameter (D). This is a disturbance factor introduced in 2002,
which allows for blasting-effects; Lorig (2007) suggested that it reduces the cohesion in
the rock mass of a slope.

The GSI can be estimated directly or it can be related and derived from another
classification approach such as Barton’s Q system, or Bieniawski’s RMR. Either way, it
is the rock mass classification component of the Hoek-Brown criterion, which is based
on a limited amount of empirical evidence, with assessment via qualitative/subjective
geological observations. Therefore a discrete representation of the fracture network is
lacking in this approach (contrasting to the approach outlined in Section 2.5.2).

Recent research has suggested variations on the GSI system. Cai and Kaiser (2007),
have shown that the basic GSI represents a peak strength that can be downgraded
using observations from the field mapping, to arrive at an estimate for the residual GSI
(GSIr) for the rock mass. This can then be used to derive Hoek-Brown parameters for
both peak and residual strengths. The calculation of GSIr incorporates a parameter
that accounts for rock-bridges; therefore this improves the ability of the GSI to describe
discontinuous rock masses. Note that a quicker but empirical expression to calculate
the GSIr from the peak GSI, was also presented by Cai and Kaiser (2007),
demonstrating that GSIr is approximately 25 to 40% of the peak GSI for hard rocks and
almost equal to the peak GSI for weak rock masses.

However, careful consideration of the discontinuous nature of the rock mass is needed
when applying the Generalised Hoek-Brown criterion. Typically the criterion is used in
continuum conditions; it is therefore applicable to strength assessment in mass
controlled-circular failure. Hoek (2007) presents the applicability of the Hoek-Brown

71
CHAPTER 2: Literature Review

criterion using the idealised transition from intact to jointed rock masses with increasing
sample size, as reproduced in Figure 2.9.

Figure 2.9: Diagram showing in what situations the Hoek-Brown criterion


for mass strength should be applied (modified after Hoek, 2007).

The Hoek-Brown criterion, coupled with the GSI system, can be considered and
automated within the Windows based freeware “RocLab” (Rocscience, 2008). RocLab
enables one to derive Generalised Hoek-Brown criterion properties, with graphical
presentation and quick/simple sensitivity analysis of input parameters.

The programs for determining factors of safety in rock slopes were initially based on
the Mohr-Coulomb shear strength criterion, as this was widely (and previously) used in
soil mechanics. Therefore many numerical modelling packages require Mohr-Coulomb
strength inputs (cohesive and frictional strength), to describe the shear strength of
intact rock, rock joints and rock masses. However, the Hoek-Brown failure criterion is
being increasingly applied, with most numerical modelling programs now able to use
Hoek-Brown among many other strength criteria. This must be contrasted though, with
data presented in Chapter 8 which proved that potentially in certain cases, the Mohr-
Coulomb failure criterion is more appropriate for representing the mass strength of a
fractured mass. This could be because the Hoek-Brown criterion was designed for
isotropic and homogeneous masses as shown in Figure 2.9; therefore as Lorig (2007)

72
CHAPTER 2: Literature Review

emphasised, the Hoek-Brown criterion is not to be applied to discontinuity-controlled


slopes.

Because the Mohr-Coulomb strength criterion is the conventional approach to intact


and mass rock strength, many numerical models are still based around Mohr-Coulomb
inputs. However, the derivation of Mohr-Coulomb parameters from a Hoek-Brown
envelope can be problematic considering the limited axis, over which the linear Mohr-
Coulomb envelope can be fitted to a curvilinear Hoek-Brown envelope. As indicated in
Figure 2.10, it is necessary to consider a range of normal stress over which the Mohr-
Coulomb failure envelope can be fitted. In the case of the Hoek-Brown approach, this
is performed via the application of the minor principal stress (σ3). Consequently, for
each rock mass situation that is under analysis in RocLab, an individual maximum σ3
(σ3max) must be defined; it is to this value that the axis is fixed and the linear Mohr-
Coulomb envelope is fitted. The use of a value of σ3max requires an a proiri assumption
on the depth of the specific failure surface on which the slope could fail.

Figure 2.10: Illustration of the method


by which Mohr-Coulomb shear
strength parameters are derived from a
curvilinear Hoek-Brown envelope.

An alternative is to use the Hoek-Brown parameters to directly simulate an equivalent


continuum within a numerical modelling package. However this eliminates the
possibilities of preliminary simple chart-based analysis, discussed in the following
section, because current slope design charts are based upon Mohr-Coulomb strength.

73
CHAPTER 2: Literature Review

Slope stability charts


Empirical chart-based methods designed to appraise slope stability, such as those
presented by Hoek and Bray (1981), can be used to determine mass strength from
back-analysis of circular failure through a mass-controlled material. Discussion on this
mode of failure was given in Section 2.2.5.

Slope stability charts are used to assess the slope mass as a continuum, through
which circular failure will occur. Therefore slope stability charts are limited in their
application to:

 weak rock masses,


 heavily fractured slopes, or
 equivalent continuums, as demonstrated in Charter 8, where mass strength
properties have been downgraded to allow the consideration of discontinuities
in a non-discrete manner.

Importantly all of these have limited cohesive strength, and mainly rely on friction to
provide mass strength.

The Hoek and Bray charts require (or can back-analyse) Mohr-Coulomb mass strength
parameters. Their ease of use and indirect approach make them a quick technique. In
addition, five different charts have been created for different groundwater conditions,
although in this respect they are limited as the influence of ground water on a shallow
circular failure cannot be captured very well. This is a result of the assumption that the
phreatic surface follows a perfect curve whereas, as discussed in Section 2.4.1, it is
more likely that rock slopes have a network of interconnecting fractures forming a semi-
saturated zone.

As discussed previously within this section, the limitations of all empirical approaches,
including slope stability charts, is that they are restricted by the scope of empirical
relations on which they were originally established. The difficulty of obtaining case
study large-scale slope failures, limits the usefulness of the charts (or indeed any
empirical-based method of analysis) as a back-analysis tool to determine rock mass
strength.

Also when using the slope stability charts as a design tool, one requires Mohr-Coulomb
mass strength parameters; following the theme of this section, such properties can be
problematic to obtain. At present there are no slope stability charts that directly take
Hoek-Brown mass strength properties, instead the only option is to derive Mohr
Coulomb strength from a Hoek-Brown envelope, which is highly dependent on σ3max, as

74
CHAPTER 2: Literature Review

discussed in the previous section, with a practical example in Section 8.3.2. Instead
perhaps one should look to derive a representative Mohr Coulomb mass strength
directly from the rock mass via numerical modelling, as discussed in the following
section.

2.5.2 Use of numerical modelling for the determination of mass


strength
The explicit incorporation of structure via discontinuum modelling may be more
accurate than attempting to mimic results by creating a downgraded equivalent
continuum. However at present it is not computationally viable to perform
geomechanical modelling on a large-scale structural model, such as a 1000m slope.
Elmo et al. (2007a) suggest that lower values of tensile strength and fracture energy
can be used in weak rock masses to reflect the influence of fractures, with either Mohr-
Coulomb or Hoek-Brown Criteria to represent the continuum or equivalent rock mass.
An equivalent mass (continuum) integrates the discontinuous nature of rock mass by
accordingly downgrading strength properties in an attempt to represent the
unfavourable strength characteristics of a specific set, or sets of unfavourably
orientated discontinuities.

This degradation of strength also occurs within empirical systems during the
classification-based approaches, which were discussed in the previous section.
However by using a numerical-based approach, the degradation of strength can be
performed in a direct way through geomechanical modelling of the DFN.

At present only two companies Itasca and Rockfield, have the relevant codes,
(UDEC/PFC and ELFEN respectively), commercially available to allow accurate
geomechanical modelling of a DFN. The following two sections discuss research from
each of these approaches, with a final section briefly outlining the subsequent
simulation of a fractured rock slope using the numerically-derived mass strength.

75
CHAPTER 2: Literature Review

FracMan-ELFEN Approach
One of the aims of this thesis is to contribute to the ‘FracMan-ELFEN’ approach, which
involves the statistical generation of a fracture network through DFN numerical
modelling software FracMan (Golder, 2008). The fracture network can then be
incorporated into ELFEN which can be used to derive a strength estimate for the rock
mass, based upon the simulated geomechanical behaviour of the rock and its
fractures. To date, the FracMan-ELFEN approach has only been applied to assess the
strength of jointed mine pillars, during the research carried out by Elmo (2006) which is
reported in Elmo et al. (2005) and Pine et al. (2006). The rock mass strength assessed
by Elmo (2006) showed encouraging similarity with empirical schemes, although the
rock mass strength that was calculated by Elmo (2006) has been shown to be
speculative, as outlined in Section 8.5.2.

In addition various other detailed aspects concerning fracture intensity have been
researched at Camborne School of Mines. For instance a relationship between the
areal 3D extent of fractures (P32) modelled by FracMan, and RMR, is reported in Ford
(2008). Also correlations have been made with rock mass behaviour and the length of
fractures within a 2D plane (P21), using FracMan and ELFEN. Pine et al. (2006)
discuss the data from Elmo (2006) which shows a critical influence of fracture
intensities (P21) on the strength of pillars; with higher P21 causing weak behaviour
especially in slender pillars. Flynn and Pine (2007) extended this research from
pillars, to demonstrate that a low P21 results in higher rock strength; however there is a
substantial scatter about this general relationship, which is due to the influence of large
fractures. Subsequently it could be suggested this relationship between the P21 and
rock mass strength can only be used when considering mass-controlled failure, where
the DFN is composed mainly of fractures which are less than a certain length (related
to the pillar dimensions). If large fractures exist within a DFN, these are likely to
contribute to structurally-controlled failure which needs a more direct (purely ELFEN)
analysis than the FracMan-ELFEN equivalent continuum approach used in Chapter 8.

Elmo et al. (2007b) provide a useful discussion on the FracMan-ELFEN research on


discontinuum behaviour and the assessment of rock-bridges within large slopes. They
introduced a relationship between P20 and P21 (total number of fractures per unit area
and total length of fractures per unit area respectively). Elmo et al. (2007b) suggest
that the P20:P21 ratio can be used as an indicator for the amount of rock bridges; for
instance low P20:P21 ratios indicate longer fractures and consequently fewer rock
bridges. Ultimately this relationship can therefore provide an indication to rock mass
integrity, with a good correlation alongside RMR.

76
CHAPTER 2: Literature Review

It can be considered that there are three parts to the FracMan-ELFEN numerical
modelling approach:

1. The initial collection of detailed geotechnical data via field mapping, sample
collection and geomechanical testing for intact rock strength; from this one can
gather the primary inputs of a model.
2. Statistical modelling of the DFN, generating a 3D fracture network via the use of
the FracMan software.
3. Finally the import and subsequent geomechanical modelling on the FracMan
fracture network, to derive a mass strength for the fractured mass, using ELFEN.

A practical application of the FracMan-ELFEN approach is performed in Chapter 8, with


parts 1 and 2 from this process gathered from Elmo (2006), who also briefly addressed
part 3. Data gathered from Middleton Mine, Derbyshire, is used to derive a mass
strength for the Hoptonwood Limestone. It was discovered that the Hoptonwood
Limestone is a relatively competent rock mass; consequently in order to fully apply the
FracMan-ELFEN approach, the compressive strength of the Hoptonwood Limestone
was degraded to give a new weaker rock mass, which retained the original Middleton
DFN. Subsequently a slope composed of new degraded rock mass is modelled as an
equivalent continuum and new mass properties are presented in Section 8.4, which are
at the limit point for a particular slope design.

An additional aspect that has been considered with the FracMan-ELFEN approach is
the rotation of the biaxial models with respect to shear strength within a DFN where
there is a dominant fracture orientation. Subsequently a change in rock mass strength
with depth is simulated. Section 8.4 presents an analysis of the strength of a rock
mass that is discretized into zones of distinct strength.

As discussed in Section 2.3.9, in some cases a 3D analysis of a slope may be


preferable; although it is only computationally feasible at present to perform
geomechanical modelling of a DFN in 2D. Elmo (2006) performed some triaxial
modelling in ELFEN in 3D, however models only had a single through-going fracture; a
full DFN would be too complex to mesh and simulate using the current single processor
system of ELFEN. Instead it is only possible to model a 3D equivalent continuum using
ELFEN, although this has not been achieved within the current duration of research,
but should be an area for further study.

77
CHAPTER 2: Literature Review

PFC-FLAC/UDEC (“Synthetic Rock Mass”) Approach


A recent advance in rock mass strength derivation published in literature, is the
Synthetic Rock Mass (SRM) approach using the various Itasca codes. The new
equivalent continuum technique has been developed by as part of the Large Open Pit
(LOP) project (CSIRO, 2008b). Mas Ivars et al. (2007) suggest that during the SRM
approach a sphere is extracted from a DFN, and is subsequently subjected to
compressive tests within either UDEC or PFC. Other than this there is no further detail
presented within the limited amount of literature, concerning the respective methods
utilized during the Itasca SRM fracture-based modelling.

Cundall (2008) suggests that the SRM approach allows one to consider the influence of
joint fabric on the strength, modulus, anisotropy and brittleness of large-scale rock
mass behaviour. Mas Ivars et al. (2007) presented a case where the strength derived
from the SRM is incorporated into a FLAC model which was used to predict three
zones of behaviour within a block caving operation; this was then calibrated with
seismic data. However a subsequent conclusion was that the SRM approach is
limited, with no consideration of pore pressures, temperature or time-dependent
deformation. Also Mas Ivars et al. (2007) state that at time of publishing, the SRM
approach had only been developed to look at the interrogation of rock mass at a scale
of 10-100m; consequently many of the micro-geological (grain-scale) features that
could have an influence on the mechanical behaviour of the mass, have not been
addressed explicitly. This is also the case with the FracMan-ELFEN approach;
although collaborative photogrammetry-FracMan research was planned upon the
small-scale behaviour of chalk (Flynn et al., 2007). This was not completed and
therefore remains an area of recommended research.

The SRM and FracMan-ELFEN approach are similar; both techniques are reliant on
the input of a DFN; the main difference is the respective numerical codes used for
geomechanical modelling. Within an introduction of the SRM approach by Read
(2007), a few important qualities of the SRM approach can be noted:

 The SRM is independent of any failure criterion;


 Calibration of the PFC model with UCS, Modulus and/or Poisson's Ratio during
an initial elastic continuum modelling stage.

Importantly both approaches result in a mass strength that represents an equivalent


continuum, which has to be then considered within a continuum-based method, as
discussed in the following section.

78
CHAPTER 2: Literature Review

Simulation of numerically-derived mass strength


FLAC3D, ELFEN or ABACUS were all suggested by Read (2007) as potential 3D
continuum and/or discontinuum modelling packages that can be used to simulate an
equivalent material determined from a SRM. The derivation of an SRM and further
large-scale equivalent continuum simulation can therefore all occur within ELFEN. The
examples of the SRM approaches presented by Mas Ivars et al. (2007) and Lorig
(2007) all require at least two modelling programmes (PFC/UDEC and FLAC/3DEC),
where either FLAC or 3DEC provide the means of considering the resulting slope mass
strength.

A potentially unique all-in one approach could be followed using ELFEN, without the
need for another continuum based code. Note that a continuum ELFEN model has not
been used to simulate rock slope mass strength during the FracMan-ELFEN approach
outlined in Chapter 8; this is due to lack of experience. Instead less complex limit
equilibrium and finite element methods have been used to analyse the mass strength
derived from the FracMan-ELFEN approach, as presented later in Section 8.4.

Prior to the final consideration of a large-slope within Chapter 8, the preceding chapters
(4 to 7) introduce an increasing range of slope case studies. To simulate the case
study slopes effectively, using a dynamic fracture-based code, the approach to
simulating a fractured slope model needs to be taken into account. Suitable
techniques of loading and slope release, based on the lessons learned from this
research, are presented within the following Chapter.

79
CHAPTER 3: Staged modelling approach

CHAPTER 3 – Development of a comprehensive


numerical model
This Chapter describes general issues critical to the development and use of dynamic
fracture-based codes. However, the context is specifically with the use of the code
ELFEN, (versions 3.8.3, 3.8.5 and 3.9.1), available at CSM during the current research
period (2005-2008).

Fundamentally, by building a fractured rock slope model, some of the processes


described in Section 2.2 are being represented during an extremely short time scale, in
a bounded environment. Within all slope models, an appropriate initial geostatic
stressed state must be achieved; this involves the application of load, which generates
internal strain energy. Within a comprehensive fracture-based code, the dynamic
response to this energy involves a change from strain to kinetic energy, the influence of
which depends on the continuum and discontinuum aspects of the model.

If a slope is created instantaneously in the model, this energy can rebound, causing
damage and artificial failure (displacement or fracturing). Consequently the simulation
of a fractured rock slope, within dynamic fracture-based codes, has to be designed so
that loading occurs in the most realistic way. This requires careful consideration of the
energy changes caused by the loading and initialisation process; with the key
modelling stages controlled, energy release rate is minimised to mimic longer-term
processes (years for mining, millions of years for natural slopes).

The ability to adopt the most realistic procedure improves with experience; therefore
this Chapter relays some of the lessons learned during the progress of this research.
An approach is proposed, to appropriately load a sensitive, fractured rock slope and
realistically observe failure. The following section presents specific issues with the
necessary slope modelling stages when using a dynamic fracture-based code;
consequently appropriate procedures are suggested to minimise the influence of kinetic
energy.

80
CHAPTER 3: Staged modelling approach

3.1 Staged modelling approaches within fracture-based codes


The benefit of a time stepping procedure within a numerical model, provides the ability
to manage the duration over which loading or unloading occurs. Importantly this also
allows one to monitor the level of kinetic energy within the model which is especially
important when considering dynamics within a fractured mass. It is essential within a
slope model to provide long enough stages within the progress of a model, to allow
equilibration of kinetic energy via damping (see Appendix A). This allows the settling of
any energy that may have been imparted by essential processes, for instance the initial
loading of the system, or excavation of the slope. If such stages are not built into the
progress of a model then an otherwise stable slope may become unstable due to the
dynamic response of the model, related to high kinetic energy release rates.

Hamman and Coulthard (2007) describe a process within their UDEC modelling of the
Sunrise Dam open pit gold mine in Western Australia, where damping allowed their
model to behave in a quasi-static manner, which is especially important within jointed
systems. For instance potential loss of shear strength and consequent inappropriate
yielding can occur if the system is not sufficiently damped, all due to transient stress-
waves (dynamics), which are generated by changes in applied loads (Hamman and
Coulthard, 2007).

3.1.1 Realistic sequence of loading and release within a slope model


Loading rates can be adjusted in an attempt to minimise energy. During early
simulations, all slope models within ELFEN were set up so they failed during a ramped
gravitational load application, due to an accumulation of stress. It is in the author’s
opinion that this is an appropriate method only when a brief review of the possible
failure mechanism is required.

In order to have a more detailed understanding of the failure process, it is important


that a steady-state situation (where kinetic energy is at/near stable), possibly on the
very limit of stability, is obtained within the slope model just prior to failure. This allows
the model to be used in a more precise way, to either:

 Back analyse a slope failure and consequent rock mass strength, or


 Further understand the mechanism that triggered failure.

81
CHAPTER 3: Staged modelling approach

The trigger mechanisms that ELFEN can presently simulate are:

 The removal of a constraint upon the slope, i.e. natural erosion or


anthropogenic excavation;
 The accumulation of damage causing time dependent degradation of the
strength of internal rock bridges;
 The build up of pore pressure within a slope, causing the reduction of shear
strength on discontinuities.

In order to simulate these processes in a kinetically stable and consequently realistic


way, a suitable loading procedure needs to be followed within a slope model. It is
proposed that the most appropriate process to construct a model is to mimic the way in
which the slope is formed in nature; this can be attempted through following a
procedure which is summarised in the steps listed below.

Proposed sequence within a slope model:


1. Load the model, establishing the appropriate geostatic regime whilst there is a
constraint upon the slope face
2. Release the discontinuities within the model
3. Remove the constraint upon the slope face at a suitable rate
4. Allowing a trigger mechanism to cause failure.

The kinetic energy imparted from each of these modelling stages should be monitored,
and used to accurately simulate a trigger process. Figure 3.1 shows the kinetic energy
during an ELFEN Delabole model in Section 6.3.2, where failure is due to a rise in the
phreatic surface. Between each of the modelling stages listed above, a steady-state
should ideally be achieved within the model during a settlement stage, as depicted in
Figure 3.1.

As discussed in the next section that details the settlement stage, these quiescent
periods, allow the interrogation and remediation of any potential artificial failures.
Consequently if failure occurs within the model before the trigger process then that
modelling stage, responsible for the dynamic response, can be lengthened to lower
kinetic energy and prevent the artificial failure.

82
CHAPTER 3: Staged modelling approach

Figure 3.1: Proposed four-step process by which fractured rock slopes should be
modelled; with component settlement periods highlighted in grey.

When using a dynamic-based code a degree of elastic-plastic displacement is


inevitable, even upon discontinuities where failure should not occur; this is related to
the dynamic response of the model. This dynamic behaviour is expressed by small
displacements that can be attributed to each of the modelling stages. Figure 3.2
presents such a behaviour, which occurred on a planar failure outlined in Section 4.5.2,
where discontinuity strength dictated a FOS of 1.1 (see Section 2.2.2). Consequently
displacement stabilises after a degree of plastic deformation.

Importantly, in such cases close to limit point limit, a state of stability is only achieved
by minimising the kinetic energy during each modelling stage. If kinetic energy is not
minimised, then inappropriate (artificial) failure can occur, resulting in instability despite
the FOS being above 1.

83
CHAPTER 3: Staged modelling approach

Figure 3.2: Dynamic behaviour within an ELFEN model, with displacement of


a discrete block due to successive modelling stages.

As discussed earlier within this section, one can assign a specific loading function to
the model. In addition to this, there are several ways by which the model can be
unloaded, in order to create the slope. In some cases there appears to be minimal
differences within models regardless of the way in which they are loaded or released.
Particular investigations of this can be found in:

 Section 4.5.2, where there is brief discussion on differences in contact pressure


of a discrete failure block due to loading and unloading approaches. However
despite this, net displacement within each method is the same.
 Section 5.2.4, which outlines a series of models where different unloading
methods were simulated, (see Figure 5.9), resulting in slightly different
damaged states within each model, but importantly similar final stresses.

However, in other cases where energy is more critical to stability, the loading and
release methods can be very influential, as found within Chapter 6. Other comparisons
have been made, as outlined in the following sections. Also there is an introduction to
each of the modelling stages, and suggestions on what is believed to be the most
appropriate conduct.

84
CHAPTER 3: Staged modelling approach

Loading
The excavation and relaxation processes within the current version of ELFEN (3.8.5),
are features that were not available in earlier codes used during this research (3.7.0
and 3.8.3). Consequently during initial modelling it was necessary to apply gravity via
either a ramped or sine curve as an attempt to minimise kinetic energy within a model
where there is no constraint on the slope face. However, since constraint techniques
are now available within ELFEN, one can use instantaneous (drop) loading for gravity.
This is because movement within the slope face and other boundaries can be
restrained, which helps to prevent dynamic effects occurring during loading.

Along with gravity loading, it is necessary to establish an in-situ stress state, which may
reflect non-gravitational (tectonic) stages of loading. Within ELFEN this is performed
within a ‘geostatic settling stage,’ which occurs at the same time as gravity loading.
The geostatic settling is assigned via the ‘geostatic state data;’ this is inputted within
the text-based file that controls the code, known as the neutral file. The following points
summarise the inputs required for the geostatic state:

1. Gravity *
2. k ratio, as discussed in Section 2.2.2.
3. ‘Zero stress horizon,’ allowing the recognition of ground surface elevation from
which the model is created.
4. Finally a load curve is assigned, which follows the same time and load-set as
that applied to the gravity load.

*In addition to the application of gravity within the geostatic state, global or local gravity
has to be applied, which is active over all surfaces within the model. Typically a load
curve is applied to the global gravity, which matches the load curve that is applied to
the application of the geostatic state.

The load curve assigned to the geostatic state data and gravity load, may influence the
dynamics within the model. Brief comparisons have been made between models
loaded via drop or ramped gravity load curves. This demonstrated that there is very
little difference between such loading approaches. In particular, within a simple planar
failure noted in Section 4.5.2, it was found that there was a difference in the kinetic
energy at the beginning of the simulation where a ramped or drop gravity load curve is
used. This led to a slight difference in excavation models, between the degree of
contact upon the discontinuities defining the planar failure, when using each of the
different loading methods. However, this difference is eliminated after the release of

85
CHAPTER 3: Staged modelling approach

discontinuities, when complete contact is gained within only excavation models (see
following section which describes the behaviour within a constraint-release model).

It is therefore beneficial, in terms of time and computational efficiency, to apply loading


via an instantaneous (drop-load) approach. In some more sensitive cases it may be
necessary to ramp load a model (one particular case is reported in Figure 6.12).

Primarily during this research, models were loaded at a comparatively slow rate in the
hope of reducing internal dynamic forces, whilst there was no constraint applied to the
slope face. With the ability to now load a model via a drop load and still achieve a
steady state, one can design the loading sequence over a significantly shorter duration.
The process, which Rockfield (2007) use in their worked example of a saturated
fractured slope, occurs in two stages with loading during the first step (via a drop load
curve), and finally discontinuities released during the second step. This process is
quick, however there is no time built into the method to allow one to check that a
steady state is achieved within the model. Consequently it is considered preferable to
allow a settlement period between these two steps.

Settlement periods and damping


The settlement periods within a numerical model are vital for an accurate and realistic
simulation of slope stability using a dynamic fracture-based code. During the
settlement periods illustrated in Figure 3.1, point damping (see Appendix A) is used to
diminish kinetic energy that was imparted as a dynamic response to one of the
modelling stages. The settlement period between modelling stages allows the cause of
fracturing during the progress of a model, to be identified. If this fracturing can be
assigned to a peak in kinetic energy that was due to a modelling stage, then the
fracturing is likely to be a dynamic effect of the modelling. This kinetic energy should
be minimised by reviewing the duration of both the modelling stage and the settlement
period:

 If the modelling stage is too short then kinetic energy may be large and artificial
failure may occur prior to the settlement stage.
 If the settlement period is not sufficiently long a settled state will not be
achieved as damping will not be active over a long enough duration (before the
next modelling stage). Subsequently any excess kinetic energy will not be
diminished and failure may occur at the beginning of the next stage.

86
CHAPTER 3: Staged modelling approach

Importantly Elmo (2006) discovered that displacement (point) damping drastically


inhibits fracturing; consequently he ensured damping was only active during the
loading stage. During the progress of this research, reported in Section 4.3, it was
found that a very specific degree of point damping is required to obtain kinematic
release of a block which is sliding on a single discontinuity. It was as a response to this
issue, that the staged process was developed. Subsequently if desired, point damping
can be removed from any critical stage, where fracturing and initiation of the resulting
failure mechanism, is simulated.

However, significant issues arise with under-damped models, during the comminution
of failure blocks from a slope. In the post-failure stages of a fractured rock slope
simulation, dynamic forces can occur even throughout the deposition of failure blocks.
This can cause further (unrealistic) failure of the slope and drastic energy effects
(bouncing) within blocks. Therefore within slope models, it has been necessary to
enable damping throughout the whole simulation.

It could be suggested that damping could be removed just at the initiation phase of
slope failure, this way no fracturing will be inhibited; however this may add to the
complexity of a simulation. Instead it is suggested that there should be more research
into the influence of dynamic-related activity due to kinetic energy, with reference to the
essential damping required within fracture-based codes. This would help to provide
guidance on acceptable levels of kinetic energy and appropriate settlement durations
(dependent upon the degree of damping). With the present situation, the development
of a staged simulation allows the examination of damping affects, and the ability to
track dynamic stability from loading to release of a fractured rock slope.

Geostatic contact initialisation (discontinuity release)


To develop a kinetically stable model, movement upon discontinuities can be fixed
during the loading period. Subsequently discontinuities can be released over a
duration which is set with respect to minimising kinetic energy, prior to the release of
the slope face.

Barla et al. (2003) present a similar practice within UDEC, where equilibrium is
achieved within their model via preventing shear along discontinuities during an
initialisation phase. Subsequently the discontinuities are then assigned strength and
deformability parameters prior to any excavation stage.

87
CHAPTER 3: Staged modelling approach

As previously stated, the release of discontinuities within a model needs to occur over
a time-scale, which adequately minimises kinetic energy. An example was given within
Rockfield (2006), where discontinuities were released within a jointed rock mass over a
period of 9s. However a subsequent simple example was given within Rockfield
(2007), where discontinuities were released within a jointed rock slope over a period of
1s. Consequently, experimentation during the progress of this research has shown
that in some cases the release of joints can be relatively quick; for instance within a
model presented later in Section 4.5.2, it was found that a settled state could be
attained with the release of the joints over a period of only 0.2s.

A settlement period is typically provided following the geostatic initialisation, as shown


in Figure 3.1. This again allows a clear definition of a steady-state prior to the next
modelling stage, which in the case of a slope model is the release/excavation of the
slope.

Release Sequence
For an accurate slope model that fails as a result of a real trigger process, there needs
to be a degree of constraint applied to the slope face during loading. This prevents the
movement that could result due to dynamic forces imparted by loading. Once a
reasonable loading/geostatic regime within a slope model is established and the
discontinuities are released, the constraint holding the slope face can be removed.
This has to occur at a slow enough rate to allow damping to diminish the dynamic
forces involved. To ensure an efficient computation time, the minimum period of
release should be found, where constraint removal is long enough to prevent kinetic
energy rising to a degree which causes dynamic-related fracturing. As with the loading
and discontinuity release rates, the only way to find an appropriate rate of release for
the slope constraint, is to use trial and error; kinetic energy should be monitored to
ensure a steady-state is established prior to the relevant trigger mechanism.

There are presently three ways of restraining a slope during the loading stage:

1. An ‘excavation’ approach, whereby a meshed object is simulated alongside the


slope. There can be a single layer or several adjacent layers of material
retaining movement of the slope. Following the loading stages the layer(s)
is/are removed to create the slope.
2. A ‘structural-constraint-release’ method, where a single structural
fixity/constraint or a series of constraints are applied to the slope face. After
loading, structural constraint(s) is/are removed to mimic formation of the slope.

88
CHAPTER 3: Staged modelling approach

3. Same as method 2 above, however instead of a structural constraint, an


‘applied displacement load’ of 0m/s is applied to the slope face. This has the
same effect as method 2 and consequently the lines to which it is applied
become fixed. However, if a zero applied displacement load is applied there is
no need to apply a structural constraint, which otherwise proved problematic in
a slope with embedded discontinuities.

Rockfield (2008, personal communication) described the use of method 3 above, in


situations where an applied displacement load is removed from a body, causing it to
move back to its original position. Instead if an applied displacement load of zero is
applied once the displacement load is deactivated, then the body remains stationary at
its current location. In the case of a slope model, this functionality is not used; instead
an applied displacement load of zero is simply used as an alternative to a structural
constraint.

During modelling, specific problems were noted with each of the methods of slope
restraint. These are listed in Table 3.1, along with suggestions and consequent
disadvantages.

Table 3.1: Methods of slope restraint within ELFEN, specific problems discovered during the use of each
method, consequent suggestions and disadvantages to each technique.

Method of slope
Problem Suggestions Disadvantages
constraint
(1) Geometry Stress concentration
Rockfield (2007, personal
object against face around tip of
communication) suggest All four methods would
to be removed/ embedded
two of four possible significantly increase
relaxed discontinuities (detail
methods (outlined in the runtime
(excavation presented within
Appendix F)
approach) Appendix F)
No failure within
Rockfield (2006, personal Structural-constraint-
discontinuity-
(2) Structural communication) suggest release approach is
controlled slopes
constraint that a ‘beam element’ was only possible where
(see Section 5.2.4
assigned to slope automatically inserted into there are no embedded
where failure was not
face the ‘mei’ file, which discontinuities within
possible in a model
prevents failure. the model.
with a tension crack).
Rockfield (2007, personal
communication) reported
Failure block tips that this does not affect the
(3) Applied- forward slightly final displacement of the Further research is
displacement load during gravity loading failure block in a situation necessary to fully
of 0 m/s assigned causing loss of where the FOS was >1. determine whether this
as constraint contact within This does not occur when influences displacement
against slope face discontinuities (see using an excavation at other limit-states.
Section 4.5.2) approach, as the
excavation layer prevents
any rotation.

89
CHAPTER 3: Staged modelling approach

From Table 3.1, it is evident that using the present version of ELFEN, potential stress
concentrations, or lengthy runtimes can be avoided by using constraint-release
method. Also an applied-displacement load must be used as the restraint within
fractured rock-slopes. However further research is perhaps necessary to prove that
the solution is not influenced by the loss of contact during loading.

If the model uses either an excavation or constraint-release method, sequential


removal of either layers or constraints can occur. This is appropriate if the trigger
mechanism is the removal of this constraint, allowing the limit point to be fully
recognised at the final, critical stage of slope release. Between each of these stages, a
steady-state should be attained during a settlement period. Within ELFEN the
relaxation of the layer or constraint, marks the duration over which removal occurs.
Like the loading, this should be again set at an appropriate duration, with respect to the
kinetic energy minimisation within the model.

If the trigger mechanism is not the removal of the constraint, then to speed up the
simulation, the constraint upon slope face can be removed in a single stage; with
relaxation over a long enough duration to ensure that kinetic energy does not impart
dynamic effects. Subsequently the slope is then subject to either progressive
deformation, a rise in the phreatic surface or other forms of loading, to bring about
failure in an accurate way.

3.1.2 Discussion
From considering the processes by which slopes are formed in nature and the aims of
modelling a slope failure, a number of possible techniques have been discussed by
which one may load, constrain and subsequently release a slope model. It appears
that the most efficient method is to assign loading via a drop load, during which a
constraint is applied (via an applied displacement load), to the slope face. However, it
is recommended that further research is required to verify that the solution of a
constraint-release model is not influenced by the rotation and consequent loss of
contact; as noted in Section 4.5.2. Also it must be noted that in some cases where
failure is particularly sensitive to kinetic energy, as outlined in Section 6.3.2, drop
loading may not be appropriate. Therefore a drop-load should perhaps be the first
method that should be implemented, but if necessary ramp-loading should be
considered.

Regardless of loading and release method, kinetic energy should be monitored during
each stage, to ensure dynamic stability prior to and following the release of
90
CHAPTER 3: Staged modelling approach

discontinuities. All these processes need to occur at a rate which minimises the
influence of kinetic energy.

Modelling during this research has demonstrated that, using the current version of
ELFEN, an applied displacement load has to be used during a constraint-release
approach within a discontinuity controlled-slope model. This is a contrast to the
method suggested by Rockfield (2006), within their continuum-based worked example
(WSX009B), where a series of structural constraints was applied to the slope face.

Finally, the trigger mechanism of slope failure needs to be clearly identified. This can
have an impact on whether there should be a staged release, or relaxation of the slope
face in a single stage. The former should occur if the trigger mechanism is
excavation/erosion; the latter is appropriate if there is an alternative trigger process.

Importantly once the slope is relaxed during a stage prior to the trigger mechanism,
using ELFEN one can analyse the strength of the rock mass, via restarting the
simulation from any desired stage. This allows the back-analysis of mass rock
strength. As discussed in Section 2.2.3, the mass strength is obviously dependent on
both the intact strength and the structure within the rock mass. A way of checking that
the model parameters are reflecting realistic intact strength is to simulate intact
strength tests, as occurs in the Synthetic Rock Mass approach discussed previously in
Section 2.5.2. This can act to verify the intact strength parameters, thus constraining
the large-scale rock mass geomechanical model. An example of model calibration, for
displacement upon a planar failure within ELFEN, is given in the following section.

91
CHAPTER 4: Case study slope 1 (planar failure)

CHAPTER 4: Case study 1 – Bench-scale discontinuity-


controlled planar failure

4.1 Introduction
This Chapter contributes to aims 1 and 2 via an example planar failure geometry that is
used to evaluate the sensitivity of the contact solution within a comprehensive fracture-
based code. Throughout Section 4.3, the influence of numerical parameters is
investigated, attempting to achieve unity with a simple FOS solution derived from a
discrete limit equilibrium analysis. Consequently numerical modelling parameters are
calibrated for the latter part of this Chapter, where a bench-scale case study is used to
test the effective stress module within ELFEN.

The bench-scale models developed during this Chapter provided familiarisation with
the complex fracture-based code, ELFEN. In total four different ELFEN models are
discussed; the inputs for these can be found in Section i of the model database, along
with the model boundaries and simulation files noted in Appendix L. In addition to
ELFEN, the Microsoft Excel-based Plane_failure is used and also RocPlane (see
Section 2.3.3).

The small scale planar failure case study reported in this Chapter, forms the basis of
understanding which was subsequently applied to the larger case study slopes
presented within following Chapters. In addition, this study has presented an
opportunity to develop a suitable kinetically stable modelling procedure as discussed in
Section 3.1; techniques from this approach have been applied to the simulations
outlined in the latter part of this Chapter.

4.2 Model geometries


Preliminary simulations with slopes 1 to 3, shown in Figure 4.1, were conducted prior to
the development of a staged modelling approach; in addition models were small and
some of the earlier models had two separate surfaces. This allowed the application of
independent properties to the main slope and failure block, as shown within Figure 4.1a
and b. Failure was detected within all models, by monitoring the vertical displacement
of an element at the top of the failure block, as illustrated in Figure 4.1c.

92
CHAPTER 4: Case study slope 1 (planar failure)

Figure 4.1: The three slope models used during the calibration of model parameters.

During Section 4.5 the planar failure model was developed, extending the boundaries
of the model to create slope 4, as illustrated in Figures 4.2a and 4.2b respectively.
This gave a response that followed limit state solutions, allowing the implementation of
groundwater in Section 4.5. Also during Section 4.5 a tension crack was introduced
creating slope 5, presented in Figure 4.2c. This was used to provide a further model to
study the behaviour of the groundwater module within ELFEN. Both slopes 4 and 5 are
a progression from slope 3 introduced above; in addition to the extension of boundaries
other mesh-based numerical parameters were altered, as can be appreciated in the
Appendix L and the material database respectively.

Figure 4.2: Extent of extended boundaries (a) that were used, with other
modifications, in ELFEN planar failure models slopes 4 (b) and 5 (c).

4.3 Calibration of modelling parameters


To study the sensitivity of ELFEN to the variation of numerical parameters,
discontinuity cohesion (c) or friction () were altered independently to control
displacement in three critical situations upon the planar failure, where:

1. Displacement occurs, i.e. the FOS < 1,


2. Displacement should not occur, but the situation is on the very limit of stability
(FOS = 1),
3. Displacement should not occur (FOS > 1).

93
CHAPTER 4: Case study slope 1 (planar failure)

The discontinuity strength to satisfy each of the three situations was determined from a
Microsoft Excel-based limit-state solution, Plane_failure (Pine, 2006b), which was
briefly introduced in Section 2.3.3.

4.3.1 Sensitivity of solution to point damping


To understand the influence of point damping (see Appendix A), different limit points
were simulated. Firstly slope 1 was used, with different degrees of point damping. For
a purely cohesive discontinuity model, stability was reached at a relatively short run
time (see Appendix G.1); consequently resulting data presented in Figure 4.3a is from
1s runtime. Figure 4.3a demonstrates that there is limited convergence upon the limit
point (FOS = 1); therefore the degree of point damping is highly influential upon a
purely cohesive surface within slope 1. For a purely frictional discontinuity (Figure
4.3b), it was found that displacement was slower; consequently displacements for all
the purely frictional models are taken from 10s runtime (see Appendix G.1).

Figure 4.3: The influence of point damping (%) upon displacement within ELFEN planar failure slope 1.
(a) Displacement at 1s runtime for a purely cohesive discontinuity; (b) displacement at 10s for a purely
frictional discontinuity.

As Appendix G.1 shows, it could be suggested that in some cases the purely frictional
models should have been run for longer than 10s, therefore the behaviour illustrated in
Figure 4.3b should perhaps be reanalysed. Despite this, Figure 4.3b illustrates that in
order for displacement when FOS <1, a degree of displacement has to be permitted at
FOS = 1. Importantly 50% damping is too high, whereas between 12.5% and 20% is
more suitable in order to achieve reasonable displacement when FOS is <1.

94
CHAPTER 4: Case study slope 1 (planar failure)

With damping shown to be more critical on a purely cohesive discontinuity, a more


comprehensive study was carried out using model 1. It was found that for the
particular model (slope 1), point damping had to be either 65 or 66% for there to be
significant displacement at a FOS of 0.98 and not at a FOS of 1.01; this is illustrated
within Figure 4.4a.

Figure 4.4b illustrates the transient behaviour discovered when point damping is 64%
and the cohesive strength is varied on a frictionless discontinuity. In particular this
shows that there can be zero displacement followed by a large displacement, even
though the cohesion had increased. No explanation for this behaviour can be
provided; instead it is suggested that there should be further investigation with a review
of the kinetic energy, and implementation of a staged modelling approach.

Figure 4.4: The influence of point damping within an ELFEN model of a purely cohesive discontinuity
in planar failure slope 1, with detection of an appropriate degree of damping

Importantly Figures 4.3 and 4.4 illustrates the potential limiting influence of point
damping. This was further demonstrated by tests with a large-scale model, slope 2,
where an exponential relationship was noted.

Within the purely cohesive simulation of slope 2, cohesion has to be set at


105 kPa for the FOS to equal 0.96, at which displacement should occur. As Figure 4.5
demonstrates, in order to get sufficient displacement at 105kPa, there needs to be
displacement damping of 3%. If damping is higher than 3% there is very minimal
displacement, and any lower then the model undergoes significant fracturing,
associated with under-damped dynamic forces.

95
CHAPTER 4: Case study slope 1 (planar failure)

Figure 4.5: Exponential


relationship within ELFEN
planar failure model 2, which
has a discontinuity strength
that dictates a FOS of <1.

Further investigation was performed into the influence of mesh size within slope 2,
which, as shown in Figure 4.6, can be significant. Figure 4.6 shows that the large-
scale simulation needs further calibration, as displacement is insignificant when
FOS <1 (discontinuity  = 54°), within the finely meshed model with the lowest degree
of damping which was simulated. In the coarsely meshed large-scale model,
displacement occurs when FOS >1 (discontinuity  = 58°); therefore unlike the small-
scale (coarse mesh model) which is also displayed, it is suggested that further
calibration of penalties* and loading scheme is required within large-scale model.

Figure 4.6: Influence of mesh


density and degree of point
damping within ELFEN planar
failure model 2. In addition
results from small-scale,
relatively coarsely meshed
model 1, is displayed.

*The normal and tangential contact penalties are discontinuity-based stiffness


parameters that are assigned within ELFEN. Along with other contact parameters, they
can prevent the penetration of surfaces; although with a relation to stiffness, the
penalties can influence displacement upon discontinuities. For a more detailed
explanation of this parameter, see Appendix A (Section A.2.3).

96
CHAPTER 4: Case study slope 1 (planar failure)

It is suggested that further research is necessary into the influence of point damping.
Also the alternative forms or damping that are available within ELFEN, such as
Rayleigh damping (see Section A.2.1 within Appendix A), should be investigated. As
discussed in Section 3.1.1, Elmo (2006) ensured that displacement (point) damping
was only active during the primary loading stage within his pillar models. This has
been tried within slope models; however without damping active, dynamic-related
issues can arise during the comminution and large-scale movement stages of a slope
failure.

In response to the difficulty experienced with calibrating a suitable degree of damping,


it could be suggested that there may be another non-calibrated numerical parameter
that is influencing the behaviour. The other properties assigned within ELFEN that may
influence the displacement upon a discontinuity are:

1. Penalties
2. Mesh size
3. Computational algorithm.

A brief review of the influence of mesh density has been conducted, as presented
previously in Figure 4.6. Prior to a more detailed investigation of mesh density, the
influence of penalties and solution algorithm is studied, as reported within the following
sections.

4.3.2 Sensitivity of solution to penalties


For a detailed review of the penalty coefficient applied within ELFEN, see Appendix A.
Importantly Klerck (2000) states that the penalty coefficient is a stiffness term and the
concept of the contact penalty can be related to a point of contact between two springs
of known normal and tangential stiffness (Pn and Pt respectively). Subsequently the
penalties are understood to be equivalent to normal and shear stiffness; therefore the
range of penalties chosen during this study attempts to encompass the range of
stiffness that can be found in literature:

 The minimum that the UDEC, (see Section 2.3.6), user manual suggests is
10GPa/m for clay infilling in tight joints in granite.
 Where there is no infill within tight joints in granites, Wines & Lilly, 2003 state
that Pn can be hundreds of GPa’s/m. Coulthard (1999, cited in Wines and Lilly,
2003) suggests a maximum of 450GPa/m for jointing in basalt.

97
CHAPTER 4: Case study slope 1 (planar failure)

It appears that appropriate values for realistic behaviour within an ELFEN model of a
planar failure, are significantly lower than those suggested above. Like damping,
penalties can have a significant influence on shear displacement, with an exponential
relationship when FOS = 1. As depicted in Figure 4.7a, there is the penalties reach a
critical value which is equivalent to half of the Young’s Modulus (E) in slope 1 (E =
20GPa in all slope models). This behaviour occurred only when conditions were at
FOS = 1. With higher discontinuity strength, the relationship is less clear, as depicted
in Figure 4.7b.

Elmo (2006) used a shear stiffness of 0.2GPa/m for joints in limestone; this is perhaps
the better estimate of stiffness for the material that was modelled within this report.
Consequently the range of penalties that were studied is as below:

 Minimum Pn = 0.3GPa/m and Pt = 0.03GPa/m


 Median: Pn = 3GPa/m and Pt = 0.3GPa/m
 Maximum Pn = 30GPa/m and Pt = 3GPa/m

A comprehensive study was conducted, using slope 1 with (note that in this case
damping was deactivated from the final modelling stage when displacement mainly
occurs). This shows that there were problems achieving the realistic behaviour upon a
purely cohesive discontinuity, within the two-surface slope 1 model, as presented by
the results presented in Appendix G.2. Consequently a single-surface model, slope 3,
illustrated in Figure 4.1c, was constructed. This provided more encouraging results
(Appendix G.3), with the correlation alongside limit states according to variation of
penalties.

Figure 4.7: (a) Critical degree of penalties (Pn = 0.1Pt) within ELFEN planar failure slope 1,
when FOS = 1 (point damping = 10%). (b) Non-uniform behaviour, discovered when
attempting to minimise displacement in a situation where FOS >1.

98
CHAPTER 4: Case study slope 1 (planar failure)

Table 4.1 presents a summary of the results presented in Appendix G.3; this shows
that realistic behaviour was exhibited upon a purely cohesive discontinuity, when
Pn = 3GPa/m. Whereas the most accurate behaviour upon a purely frictional
discontinuity, was obtained when Pn = 0.3GPa/m. Additionally, in all simulations where
the discontinuity was purely frictional, failure occurred when FOS = 1, which was not
the case with a purely cohesive discontinuity.

Table 4.1: Results from simulations on planar failure slope 3, demonstrating the sensitivity of the
solution to variation in penalties assigned to the single discontinuity defining the discrete failure block.

Normal Tangential Displacement (m) at:


Discontinuity strength dictating
Penalty Penalty
FOS* FOS <1 FOS = 1 FOS >1
(GPa/m) (GPa/m)
0.3 0.03 Cohesion 0 0 0
3 0.3 (c = 33, 36.5 and 40kPa 2 0 0
30 3 FOS = 0.9, 1 and 1.1 respectively) 2 2 2
0.3 0.03 Friction 2 2 0
3 0.3 (c = 54°, 56° and 58° 2 2 2
30 3 FOS = 0.9, 1 and 1.1 respectively) 2 2 2

* The FOS for the respective purely cohesive or frictional conditions, are derived from
Plane_failure (Pine, 2006b).

4.3.3 Sensitivity of solution to further modifications


Following meetings with Rockfield on 11th and 12th July 2006 modifications were
suggested to give a better control on the simulations, removing the sensitivity to
penalty variation and increase the likelihood of a convergence at a realistic solution:

1. Addition of ‘option 166’ facet to facet contact (detailed in Appendix A)


2. Decrease critical time-step (Tcr) to 0.5 – 0.6
3. Application of point damping throughout the whole model and during all stages
4. Decrease contact damping from 30% to 10%.

The above modifications were made resulting in a significantly different behaviour.


Firstly, complete failure could not be achieved upon a purely cohesive discontinuity, as
illustrated within the time-displacement plots in Appendix H. The modifications did
slightly decrease the sensitivity of the solution to penalties upon a purely frictional
discontinuity; although with addition of the suggested amendments, failure became
non-uniform at higher penalties upon a purely frictional discontinuity (see Appendix H).

When the slope models discussed throughout this section were reviewed, it was
decided that amendments to the model design and the implementation of a more
suitable modelling approach should be investigated. The following section presents
99
CHAPTER 4: Case study slope 1 (planar failure)

data from a model where such alterations were made; this resulted in a more
successful model, slope 4, which conforms to the solution given by limit-equilibrium
analysis.

4.4 Modified slope, version 4


As discussed, this section details the results from slope 4. This model incorporates
some of the modifications suggested in Section 4.3.3, a newer version of ELFEN,
slightly different penalties and a suitable sequence of loading and release, as can be
appreciated within the model database.

Only a brief study was completed on slope 4, interrogating the influence of numerical
parameters. To summarise, the following sensitivities can be noted:

 Mesh density along failure plane – no kinematic release* upon a surface


where FOS is <1, within a model with a coarse uniform mesh of 2m. Failure is
permitted when the mesh density is reduced (assigning the discontinuity a
mesh density of 0.5m).
 Penalties – there is no kinematic release on a discontinuity where FOS is <1
(due to  being 54°), when the penalties are at 200MPa/m and 20MPa/m (Pn
and Pt respectively). However the model conforms to limit state theory, as
illustrated in Figure 4.8, when the penalties are reduced to 20MPa/m and
2MPa/m (Pn and Pt respectively), which is 10% of the range suggested by
Rockfield.

*No kinematic release means that there is a degree of displacement (as expected in a
dynamic model), however the model stabilises with the failure block remaining hung-up
with a net displacement of up to 6cm.

Figure 4.8: Vertical displacement plot of ELFEN planar failure slope 4 model;
note that where frictional strength is stated, cohesion is zero and vice-versa.

100
CHAPTER 4: Case study slope 1 (planar failure)

With the more appropriate behaviour being simulated by this model, as illustrated in
Figure 4.8, the influence of groundwater could be analysed with more confidence. This
gave the chance to study the newly implemented effective stress module in ELFEN,
checking behaviour against known limit states, as introduced in the following section.

4.5 Introduction of a phreatic surface and implications for


stability
This section introduces the simulations within two planar failure models, slopes 4 and
5. Slope 5 has a tension crack as a further addition, to test the groundwater module
within ELFEN.

4.5.1 Model without tension crack (planar failure slope 4)


A phreatic surface was introduced into the slope 4. This showed ELFEN to
successfully decrease the normal stress on the discontinuity relative to elevated pore
pressure, resulting in failure.

Note that RocPlane (Rocscience, 2008) had to be used for the limit-state calculations
within a wet slope, as opposed to the Microsoft Excel-based Plane_failure (Pine,
2006b) which has been used in all previous and following limit-equilibrium analyses
within this Chapter. This is because the input of the groundwater within Plane_failure
is based on the depth of water within a tension crack, and there is no tension crack
within slope 4.

It is problematic when analysing the pore pressure distribution within a model without a
tension crack. It could be considered that the peak pressure should be at half way
along the discontinuity, as illustrated in Figure 4.9a. This is due to the seepage and
subsequent dissipation of pore pressure where the discontinuity day-lights in the slope
face. To investigate further, an analysis was also performed where the peak pressure
is assigned to the base of the discontinuity, as shown in Figure 4.9b.

As presented in Table 4.2, there was little difference between the FOS from either of
these pressure distributions. However, as the data suggests, the behaviour in ELFEN
matched the RocPlane analysis best when the peak pressure is applied to a mid height
(specifically in the case where  = 58°, zero cohesive strength and 30% groundwater).
Further testing of limit points, where there is a difference between pore pressure

101
CHAPTER 4: Case study slope 1 (planar failure)

distributions (shown in Figures 4.9a and 4.9b), would be necessary to prove this
suggestion.

Figure 4.9: Pore pressure distributions within the two fully saturated
RocPlane (Rocscience, 2008) models, which investigate the
influence of peak pressure assignment in a plane failure analysis.

As illustrated in Figure 4.10 and Table 4.2, the behaviour within ELFEN matched the
solution predicted by limit equilibrium analysis where discontinuity  = 58° (and c = 0),
with failure occurring when groundwater rose to saturate the discontinuity by 50%. In
addition to this critical point, another limit-state was successfully simulated where there
is a discontinuity which has a mixture of cohesive and frictional discontinuity strength
with a FOS of 1.3 when 50% saturated, and a FOS of 0.8 with 100% saturation. A
summary of results from the ELFEN simulations is presented in Table 4.2, (where
failure occurred due to a rise in groundwater, the entry is highlighted in bold).

Figure 4.10: Influence of groundwater within ELFEN planar failure slope 4 model, where
discontinuity  = 58° (and c = 0). Average pore pressure is contoured in N/m and the
2

phreatic surface is also indicated.

102
CHAPTER 4: Case study slope 1 (planar failure)

Table 4.2: Analysis of planar failure slope 4, proving ELFEN to conform to simple limit states predicted by
RocPlane. (Comparisons are also included between different locations of peak pressure in RocPlane, as
illustrated in Figure 4.9).

Conditions on the single discontinuity


FOS predicted by:
defining the discrete failure block:


% water filling height RocPlane –peak pore pressure at:
c (kPa) of tension crack base of ELFEN
mid-height
(phreatic surface) discontinuity
54° 0 0 (dry) 0.93 <1
56° 0 0 (dry) 1.00 >1*
58° 0 0 (dry) 1.08 >1
58° 0 30 1.02 0.96 >1
58° 0 50 0.92 0.75 <1
0 33 0 (dry) 0.91 <1
0 36.5 0 (dry) 1.01 >1**
0 40 0 (dry) 1.11 >1**
56° 16 50 1.29 1.13 >1**
56° 16 100 0.83 0.00 <1

* 5.5 cm vertical displacement and then the model stabilises (see Figure 4.8).
** 1.6 cm vertical displacement and then the model stabilises (see Figure 4.8).

From Table 4.2 one can see that some vertical displacement occurs in all simulations.
Where this displacement is minimal and a stable state is reached, as illustrated by the
high c and  (FOS = 1.1) cases in Figure 4.8, it has been assumed that ELFEN has
predicted a FOS 1.

As outlined in Section 3.1.1, this displacement is due to the dynamic nature of ELFEN.
It must also be noted that in order to achieve a stable state in situations which are so
close to the limit of equilibrium (FOS = 1), a staged slope modelling procedure had to
be followed (as discussed previously in Chapter 3).

The aim of studying the influence of groundwater within planar failure slope 4 was to
validate the groundwater module within ELFEN, prior to the more detailed fracture-
based groundwater simulations within Chapters 5 and 6. Closer points of study are
recommended to further analyse the effectiveness of the groundwater component
within ELFEN. During this research, a set of simulations was conducted on a slightly
different model, with a tension crack. This was to provide further validation of the
groundwater module, as detailed following section.

103
CHAPTER 4: Case study slope 1 (planar failure)

4.5.2 Model with tension crack (planar failure slope 5)


The geometry of planar failure slope 4 was modified, adding a tension crack to create
slope 5 (see Figure 4.2c). During preliminary simulations it was identified that there
were differences in the contact pressure within the tension crack, depending on
whether the model was loaded by a ramp or drop load curve and released via an
excavation or constraint release. From the beginning of the simulation, a realistic
degree of contact can only consistently be ensured within a ramp-loaded excavation
model of slope 5.

Subsequent to this research, Rockfield (2007, personal communication) investigated


the differences noted within the models. As reported in Section 3.1.1, the failure block
tips forward slightly when using a constraint-release method, causing loss of contact
within the tension crack which does not occur within an excavation model. Regardless
of this it was found by Rockfield (2007, personal communication) that this does not
affect the final displacement of the failure block in a situation where the FOS was >1.
However, it is suggested that further research is required to fully determine whether
this influences displacement at other limit-states.

During this research further investigations could not be carried out due to time
constraints. Consequently a ramp-loaded excavation model was used within the
following study, as it is clear that a consistent degree of contact can be ensured
throughout the whole simulation. Another issue, contrasting to the study outlined in
Section 4.5.1, is the fact that in this case the Microsoft Excel-based Plane_failure
(Pine, 2006b) was used to compute limit states.

Table 4.3 lists the results from primary groundwater simulations using planar failure
slope 5 (again entries in bold, indicate where failure has/has not occurred due to a
rise/fall in pore pressure relative to the phreatic surface). From this, one can
appreciate that reasonable behaviour was obtained within slope 5 when the
discontinuity had purely frictional strength. When the frictional strength was decreased
and a component of cohesive strength was applied, the limit state could not be
obtained. It was discovered that the phreatic surface needed to be modified in this
case, as the frictional strength on the discontinuity was at the limit point and therefore
the solution was highly dependent on the pore pressure acting upon the discontinuity.
The following part of this section outlines how the distribution of pore pressure had to
be altered within slope 5, to attain a realistic limit-state solution within ELFEN.

104
CHAPTER 4: Case study slope 1 (planar failure)

Table 4.3: Analysis of planar failure slope 5, with failure occurring in ELFEN when Plane_failure
predicts a FOS >1, due to a simplified phreatic surface.

Conditions on the single discontinuity defining the


discrete failure block: FOS predicted by:


% water filling height of tension
c (kPa) Plane_failure ELFEN
crack (phreatic surface)
54° 0 0 (dry) 0.9 <1
56° 0 0 (dry) 1.0 <1
58° 0 0 (dry) 1.0 >1*
58° 0 50 0.8 <1
0 55 0 (dry) 1.1 >1**
0 45 0 (dry) 0.9 <1
56° 16 0 (dry) 1.3 >1**
56° 16 50 1.1 <1
56° 20 50 1.2 <1
56° 30 50 1.4 >1**

* 9 cm of net vertical displacement and then the model stabilises;


** 2 cm of net vertical displacement and then the model stabilises.

An animation of the failure (when discontinuity frictional strength is 58°), due to a rise in
groundwater, is included on the appended CD-rom as the video file “Planar failure
slope 5.” This demonstrates a small degree of primary (dynamic-related) movement
following the excavation of the slope, which stabilises; following this, failure is triggered
by a rise in pore pressure which triggers kinematic release.

Modification of phreatic surface


Table 4.3 presents the results from models where a simplified phreatic surface was
used in ELFEN, as shown in Figure 4.11a. Using this surface, failure occurred even
though the conditions dictated a FOS >1 (as shown in Table 4.3). However, when the
water table is modified so that there is less pore pressure acting upon the discontinuity,
failure is inhibited as illustrated in Figure 4.11b.

Pore pressure was reduced along the discontinuity within slope 5 by lowering the
phreatic surface, in an attempt to correlate the distribution of uplift forces from
groundwater, between ELFEN and the limit equilibrium approach used. As illustrated
in Figure 4.11c, both Plane_failure and RocPlane assume a triangular pressure
distribution along the discontinuities, with peak pressure at the base of the tension
crack as shown in Figure 4.11c, (the location of this peak pressure can be varied in
RocPlane).

Ideally to match pore pressure distribution in ELFEN with that assigned within the limit
equilibrium approach, the phreatic surface within the ELFEN model would follow a
curved path without touching the slope face. This would reduce the height of the

105
CHAPTER 4: Case study slope 1 (planar failure)

phreatic surface above the discontinuity, subsequently decreasing the pore pressure
and resulting uplift forces within the discontinuity. However for this to occur, mesh
density would have to be very fine which would lengthen run times. Instead these
simulations only aimed to give a approximate representation, providing a quick
validation on the behaviour of the groundwater version of ELFEN.

Figure 4.11: (a) Failure when FOS is 1.1 (discontinuity c = 12 kPa,  = 56°), due to
simplified phreatic surface. (b) Prevention of failure by modification of phreatic surface, in
an attempt to achieve similar pressure distribution as in limit state analysis (c).

As indicated within Table 4.4, ELFEN conformed to the limit states predicted by the
Plane_failure (Pine, 2006b); although failure did not occur until the tension crack was
80% saturated, which matched a FOS of 0.8. Consequently the phreatic surface would
again have to be modified to increase the pore pressure acting within the discontinuity.
As previously discussed, a finer mesh density is required, to allow the simulation of a
more accurate curved phreatic surface.

Table 4.4: Analysis of planar failure slope 5, where the shape of the phreatic surface has been
modified in the ELFEN models, as indicated by 4.11b.

Conditions on the single discontinuity defining the


discrete failure block: FOS predicted by:


% water filling height of tension
c (kPa) Plane_failure ELFEN
crack (phreatic surface)
56° 12 30 1.1 >1*
56° 12 50 1.0 >1**
56° 12 70 0.9 >1***
56° 12 80 0.8 <1

* 2.25 cm of net vertical displacement and then the model stabilises.


** 2.5 cm of net vertical displacement and then the model stabilises, see Figure 3.2.
*** 3.6 cm of net vertical displacement and then the model stabilises.

106
CHAPTER 4: Case study slope 1 (planar failure)

4.6 Summary and conclusions


During the early part of this Chapter, numerous simulations were conducted on a series
of simple unsaturated models (planar failure slopes 1 to 3). Calibration followed the
aim of correlating the solution predicted by ELFEN with a series of known limit states,
analysed by a limit equilibrium approach. The results demonstrated that ELFEN is
particularly sensitive to the degree of point damping assigned to the model, and also
the penalties assigned to the discontinuity.

Rockfield (2006, personal communication) suggested various modifications to improve


the solution within ELFEN, however at the time none of the modifications proved
successful. In addition it was also noted that other parameters (fracture energy and the
fracture opening flag – see Appendix H), which are not directly linked to the
discontinuity strength, also influenced the net displacement within planar failure
slope 3.

It is suggested that the boundaries within slopes 1 to 3, are too close, and
consequently may be affecting the dynamics within the model. Also at that time an
appropriate loading and release sequence, (see Section 3.1.1), had not been
developed. Both these aspects may have contributed to the difficulties reported within
Section 4.3. More suitable boundaries and the implementation of a staged modelling
approach in slope 4, allowed correlation with limit state theory, as outlined in
Section 4.4.

Subsequently, groundwater was introduced demonstrating that reasonable behaviour


could be achieved within slope 4, with a simplified phreatic surface. However, upon the
development of a model with a tension crack (slope 5), the simplified phreatic surface
that was applied to slope 4 was inadequate in situations where the pore pressure upon
the discontinuity is critical. This has illustrated the importance of matching resulting
uplift forces due to pore pressure, between ELFEN and the limit equilibrium method
being used for comparison. In some cases it may be necessary to use a finer mesh
density to improve the accuracy of the phreatic surface within simulations.

107
CHAPTER 5: Case study (chalk) slopes 2 and 3

CHAPTER 5: Case studies 2 and 3 – modelling of mixed-


mass and discontinuity-controlled planar failure at
bench-scale

5.1 Introduction
This Chapter reviews two small-scale mixed discontinuity-mass controlled failures,
contributing to both aims 1 and 2, through achieving objectives 3, 4 and 5. Section 5.2
presents the second case study slope within this thesis, which is the failure of a dry
chalk cliff at Joss Bay on the Isle of Thanet in northeast Kent, England. Within this
thesis, this model is termed the ‘Hutchinson Joss Bay’ failure, after J. N. Hutchinson
originally characterised the failure in 1970.

The step-path chalk failure, (case study slope 3 within this thesis), is similar to the
model of the Joss Bay in that it is also a model of a chalk cliff; however it explores the
potential progressive failure within an example cliff section. The limit point of failure is
found within one of the step-path models presented, to allow investigation of failure and
subsequent kinematic release as a result of a rise in the phreatic surface.

In total five different numerical modelling programs, (see Section 2.3 for detail on
each), were used during the research outlined in this Chapter. Firstly three different
limit equilibrium methods were used as outlined in Section 5.2.2; SLIDE was firstly
used, followed by Plane_failure and RocPlane. A Phase2 was then developed, and
finally an ELFEN model. Seven different ELFEN models are used during the analysis
of the Hutchinson Joss Bay failure; the numerical parameters for each are given in
Section ii of the model database, with detail on the geometry and data files given in
Appendix L. Finally several ELFEN chalk step-path models were developed in
Section 5.3; all of these used the same numerical parameters, as shown in Section iii
of the model database.

5.2 Case Study 2 – ‘Hutchinson Joss Bay-type’ chalk cliff failure


The following sections review previous research that has been completed on the chalk
cliff failure at Joss Bay and the data that can be derived from these independent
studies, which form the basis of the model created during this research. Subsequently
the findings are reported from a range of numerical models, in order of increasing
complexity, with a final discussion on the data obtained from each of these analyses.

108
CHAPTER 5: Case study (chalk) slopes 2 and 3

5.2.1 Model Development


Hutchinson (1970 cited in Hoek and Bray, 1981) first studied the undercutting and
resulting failure of the cliff section, using samples from the cliff face to determine intact
strength. Since this first work, Hoek and Bray (1981) have re-investigated the failure,
using limit equilibrium theory to provide further estimates of intact strength. Finally
Mortimore et al. (2004a) referred to the Joss Bay failure, presenting more detail upon
the geological units involved. Some of this detail, along with information from
Lawrence (2006), is illustrated in Figure 5.1.

Figure 5.1: Geometry and geology of the Joss Bay model.

Duperret et al. (2004) noted that the Seaford Chalk formation contains bands of large
flints but on the whole is more homogeneous than the Newhaven Chalk, which contain
numerous marl seams. Mortimore et al. (2004b) suggested that with unique physical
properties, each Chalk formation can potentially characterise a particular failure
mechanism. Duperret et al. (2004) suggests four different types of failures, one of
these is ‘sliding failures’ involving two superimposed chalk units, leaving a scar where
there is a change in slope profile.

No specific geotechnical properties for the Seaford or Newhaven Chalk units can be
found within published literature. Therefore, following on from the past analyses
carried out by Hoek and Bray (1981), homogeneity is assumed for the Hutchinson Joss
Bay model, comprising a single geological unit. The intact strength that Hutchinson
(1970 cited in Hoek and Bray, 1981) obtained from laboratory testing of samples is
listed in Table 5.1; in addition the reinterpreted higher intact strength values from Hoek
and Bray (1981) are also presented.

109
CHAPTER 5: Case study (chalk) slopes 2 and 3

Table 5.1: Peak mass strengths for the Chalk at Joss Bay, reported by both Hutchinson (1970
cited in Hoek and Bray, 1981) and Hoek and Bray (1981).

Intact strength
Data Source Cohesion:
Friction (°)
(tonnes/m2) kN/m2 = kPa
Hutchinson (1970 cited in
13.3 133 42
Hoek and Bray, 1981)
Hoek and Bray (1981) 2.64 26.4 50

The intact strength reported by Hoek and Bray (1981) is from a reinterpretation of the
failure using the equation that gives the inclination of a critical failure plane, stated in
Chapter 2 (Equation 2.1). Hoek and Bray (1981) suggest that the strength values from
their interpretation are significantly higher due to the roughness, not originally being
accounted for by Hutchinson (1970 cited in Hoek and Bray, 1981). For this reason it is
the latter rock strength properties, derived by Hoek and Bray (1981), which are used
within the numerical models presented in this Chapter.

Importantly Hutchinson (1970 cited in Hoek and Bray, 1981) reported that:

1. Failure occurred under dry conditions


2. There are two major almost vertical joint sets, one of these strikes parallel to
the cliff face, and therefore could form the plane of weakness along which the
tension crack formed.
3. Also bedding is within 1° of the horizontal.

As previously stated, Mortimore et al. (2004a) performed further investigations of the


structural geology of the Joss Bay, however they do not report on any significant
differences between the two units presented within Figure 5.1. Instead suggestion was
given by Lawrence (2006, personal communication) that:

 Seaford Chalk is slightly stronger, having larger vertically aligned


discontinuities
 Newhaven Chalk has a steeply inclined conjugate set, having the potential to
form block failures

The blocky failures suggested, would need to be investigated using a 3D model.


However, for the purpose of this investigation, a simple 2D analysis is performed using
a variety of modelling methods.

110
CHAPTER 5: Case study (chalk) slopes 2 and 3

The timing of tension crack development is a difficult issue; it could be suggested that
the tension crack was a pre-existing feature. Hutchinson (1972, cited in Williams et al.,
2004) notes that failure of chalk cliffs is often preceded by the opening of a joint-
controlled tension crack, extending downwards as the undercutting notch increases in
depth due to marine erosion. The progress of the cliff failure at Joss Bay is briefly
outlined by Mortimore et al. (2004a), who suggest that the tension crack developed
‘predominantly,’ with the inclined failure surface occurring as a result of intact material
failure following a complex network of minor joints. Therefore the principal model
simulates the Joss Bay failure with a pre-existing tension crack.

However, Sections 5.2.3 and 5.2.4 both introduce a model where the tension crack is
not pre-existing, using numerical modelling to simulate the progression of shear failure
upwards from the toe of the slope, to a point where the tensile failure can break to
surface. Hutchinson (1970 cited in Hoek and Bray, 1981) illustrates that the critical
depth which this notch must reach, for the respective section, is 0.5m as presented in
Figure 5.1. Hoek and Bray, (1981) also went further to suggest the depth that the
notch would have to reach for the section to fail again. This critical depth of the notch
has not been investigated as part of this research; instead an accurate representation
of the failure mechanism has been developed.

Finally, as presented in Figure 5.1, the geological contact between the two Chalk
formations involved is close to the termination of the shear plane (base of the tension
crack). Mortimore et al. (2004a) also suggest that a potential factor controlling the
shape of the failure plane could be the change in lithology within the cliff at Joss bay.
In order to analyse this further, more geotechnical detail would be required on the
specific Chalk formations; this is an area of recommended research.

The Hutchinson Joss Bay failure has been modelling using the three methods listed
below.

1. Various limit equilibrium solutions: an Microsoft© Excel-based approach known


as Plane_failure (Pine, 2006b), and also RocPlane and SLIDE, (for detail on
each method see Section 2.3.3);
2. FEM (Phase2); and
3. A more comprehensive fracture-based method (ELFEN).

Each of these methods were introduced in Section 2.3; importantly they range in their
complexity with the limit equilibrium solutions being the simplest; SLIDE allows a
search and FOS derivation (see Section 2.3.3) for an unspecified non-circular failure
111
CHAPTER 5: Case study (chalk) slopes 2 and 3

surfaces. Phase2 allows the creation of the slope via staged excavation process, with
investigation of mass strength using the SSR method. Finally the most
comprehensive, ELFEN, provides a detailed analysis of strength and strain within the
rock mass. Also using ELFEN a sequence of excavation is followed, to progressively
unload the slope, as discussed in Chapter 3, replicating the inferred creation of the cliff
face from successive erosion and failures. The results from these analyses are
presented in the following sections.

5.2.2 Results from limit equilibrium analysis


A non-circular SLIDE model provides the capability to accurately represent the
geometry of the mechanism, illustrated in Figures 5.1 and 5.2. Within this section the
data from a SLIDE model is firstly reported, followed by a review of results from other
discontinuity-based limit equilibrium analyses.

SLIDE Analysis
A non-circular method path search was conducted within SLIDE; a ‘block search’
technique is the alternative to the path search, however this appears to be more
appropriate for active-passive wedge problems (see Section 2.2.5). In the case of the
Hutchinson Joss Bay model, there was little difference between the results derived by
back-analysis using a block or path search method, based on the c and  values
shown in Table 5.2.

SLIDE automatically predicts that the most critical failure surface, with the lowest FOS,
is higher within the slope than the failure surface that is presented within Hoek and
Bray (1981). To achieve results that were comparable between each slice-based
analysis method, a failure surface was chosen within SLIDE that most closely
represented the actual failure surface, as presented in Figure 5.2a.

The material properties used within SLIDE are shown within Table 5.2. Note that the
unit weight of chalk, used in the limit equilibrium analyses, was obtained from a density
that was suggested by Lawrence (2006), as a typical value for chalk. The density
suggested by Lawrence (2006) was 1700 kg/m3; equivalent to a unit weight () of
17 kN/m3.

112
CHAPTER 5: Case study (chalk) slopes 2 and 3

Table 5.2: Strength inputs for limit equilibrium analysis of mass strength.

Parameter Value/Input
3
Unit weight 17 kN/m
Strength type Mohr Coulomb
Cohesion 26.4 kPa
Friction 50°

Figure 5.2: (a) Failure surface chosen for analysis of the


Hutchinson Joss Bay model in SLIDE (Rocscience, 2008). (b)
FOS from non-slice limit equilibrium method RocPlane, which
can only analyse a simplified geometry.

As shown in Table 5.3, the average FOS from the different slice-based methods in
SLIDE, is 1.05. It is recommended that further research is required into the
performance of each of the slice-based methods within SLIDE, when analysing a non-
circular analysis. As outlined in Appendix C, the Bishop and Ordinary/Fellenius
methods are suggested to be not applicable to the analysis of non-circular methods,
which perhaps is why both these methods are giving a high FOS (see Table 5.3), for
the specific failure surface presented in Figure 5.2a. The Janbu methods give a lower
FOS, and the Lowe/Karafiath proposes a FOS that is just above 1. By averaging all of
the methods, a FOS is gained and presented in Table 5.3, which is similar to the
discrete analyses that are outlined in the following text.

113
CHAPTER 5: Case study (chalk) slopes 2 and 3

Plane_Failure and RocPlane Analysis


In addition to SLIDE, two other discontinuity-based limit equilibrium solutions were
derived, using RocPlane (Rocscience, 2008) and the Microsoft© Excel-based
Plane_Failure (Pine, 2006b). Simplified geometries had to be constructed from the
dimensions suggested by Mortimore et al. (2004a) as presented in Figure 5.2b,
because both methods cannot analyse a model with a change in angle within the slope
face, or a notch at the base. Plane_Failure and RocPlane gave similar factors of
safety, averaging to a value of 1.09; this slightly higher FOS could be a consequence of
the simplified geometry. Further testing is recommended, with a numerical model of
the simplified geometry.

The three limit equilibrium methods have demonstrated similar results, with the non-
circular slice SLIDE model predicting the lowest FOS. Importantly all methods, have
shown that the mass strength predicted by Hoek and Bray (1981) is close to the limit
for failure.

A summary of the results from the limit equilibrium analyses is given in the Table 5.3:

Table 5.3: Summary of results from different limit equilibrium analyses, using the properties and
geometries presented in Table 5.2 and Figure 5.2.

Analysis Method Factor of Safety


SLIDE – Bishop (simplified) 1.23
SLIDE – Ordinary/Fellenius 1.25
SLIDE – Janbu (simplified) 0.82
SLIDE – Janbu (corrected) 0.85
SLIDE – Lowe-Karafiath 1.08
SLIDE AVERAGE 1.05
Plane_failure 1.06
RocPlane 1.12
Plane_failure/RocPlane AVERAGE 1.09

114
CHAPTER 5: Case study (chalk) slopes 2 and 3

5.2.3 Results from hybrid FEM/BEM (Phase2)


Various Phase2 models were created of the Hutchinson Joss Bay failure, primarily with
the aim of building a model to compare with the ELFEN simulations which follow. The
SSR method, introduced in Section 2.3.4, was used, demonstrating that the mass
strength identified by Hoek and Bray (1981), could be slightly lower.

The more comprehensive methods of analysis detailed in this, and the following
section, demand numerical parameters that are not required within the limit equilibrium
techniques described in the previous section. The following section details the
derivation of tensile strength (σt), which is one of the important parameters required
within Phase2.

Tensile Strength
Lawrence (2006) suggested an intact tensile strength of 1.5MPa, from testing on chalk
performed by the University of Brighton. However, mass strength properties are
required for the finite element mesh, therefore the intact strength has to be
downgraded to allow consideration of the micro-structure and to consequently
represent the rock mass. This can either be downgraded via the use of an empirical
relation, back analysis of a failure or direct modelling of the discontinuous mass, as is
reported in Chapter 8.3. In this Chapter, the intact tensile strength is downgraded
using an empirical-based approach, RocLab (2008).

RocLab was used to back analyse a suitable mass tensile strength (σtmass). The
method followed was to select a suitable compressive strength (σci), mi and σ3max
(described in Section 2.5.1), and vary the GSI (see Section 2.5.1) until a c and  were
gained that are close to those suggested by Hoek and Bray (1981), presented in
Table 5.1. The σ3max can be assumed to be close to the σv, which is calculated from 
and the average depth of the failure plane (h), through the following equation:

σv = h [5.1]

In this case the average depth to the failure plane is approximately 4m; consequently:

v = 17 kN/m3  4 m = 68 kN/m 2
v = 68 kPa = 0.07 MPa

115
CHAPTER 5: Case study (chalk) slopes 2 and 3

From field observations of the chalk units within the cliffs near Brighton, a GSI of
between 30 and 40 would be appropriate; RocLab (2008) suggests a σc of between 5
and 25MPa for chalk, and an mi of 7 2. These variations have been taken into
account within Table 5.4, demonstrating that the closest match to the c and 
suggested by Hoek and Bray (1980) is obtained using:

 An intermediate value of σci (case 3);


 a low GSI (case 3);
 an mi just below the range suggested within RocLab, as otherwise a high mi
results in perhaps an unrealistically low σtmass (case 5); and,
 a σ3max which is slightly lower than that which was calculated (case 6).

Table 5.4: Derivation of mass tensile strength for chalk using RocLab (Rocscience, 2008) and the Mohr-
Coulomb mass strength, suggested by Hoek and Bray (1983).

RocLab input RocLab output


 (°)
Case σ3max σtmass
σc GSI mi c (kPa)
(kPa) (kPa)
1 5 30 7 70 25 44 4
2 25 30 7 70 64 55 18
3 10 30 7 70 36 49 7
4 10 40 7 70 55 52 15
5 10 30 9 70 36 51 6
6 10 30 4 30 30 49 13

In addition to the parameters discussed within this section, a suitable k ratio is required
within Phase2. This can be calculated to be 0.32 using Equation 2.3, from the
Poisson’s ratio of 0.24 suggested by Lawrence (2006). Also within the Hutchinson
Joss Bay model, k1 = k2 (i.e. there is no tectonic stress). To summarise all of the
inputs, the parameters for the Phase2 model are listed in Table 5.5.

116
CHAPTER 5: Case study (chalk) slopes 2 and 3

Table 5.5: List of material properties, used within the Phase2 simulation of the Hutchinson Joss Bay model.
(Where appropriate the source, from which the strength parameter has been obtained, is stated).

Parameter Value/Entry Source


Initial element loading Field Stress & Body Force -

Stress
0.017 MN/m3

Field
Unit weight Lawrence (2006)
k ratio (in-plane) 0.32 Calculated from
k ratio (out-of-plane) 0.32 Poisson’s Ratio
Elastic type Isotropic -
properties
Elastic

Young's modulus 1 GPa = 1e9 N/m2 Lawrence (2006)


Mass strength properties

Poisson's ratio 0.24 Lawrence (2006)


Failure criterion Mohr Coulomb -
Material type Plastic -
Strength parameters

Tensile strength 13 kPa RocLab-derived


Dilation angle 5° -
Hoek and Bray
Peak friction angle 50°
(1981)
Residual friction angle 50° -
Hoek and Bray
Peak cohesion 0.0264 MPa
(1981)
Residual cohesion 0.0264 MPa -
(Mohr-Coulomb)

Tensile strength 0 MPa -


Joint strength parameters

Slip criterion

Cohesion 0 MPa -

Friction angle 30° -

100 times below


Normal stiffness 1000 MPa/m Phase2 default, to
Stiffness match the degree
Shear stiffness 100 MPa/m of penalties used
in ELFEN

Shear strength reduction


During sensitivity analyses, it was discovered that the residual strength had a
significant influence on the consequent SRF (see Section 2.3.4) computed during a
SSR simulation. Consequently as listed in Table 5.5, the residual strength remained
the same as the peak strength to ensure an elasto-plastic response.

Two models were simulated, one with a tension crack and one without. Within the
tension crack model, either an ‘open’ or ‘closed’ condition has to be assigned to the
ends of the joint. If open, movement is permitted at the joint end, having two nodes
inserted into the FEM; whereas when closed only one node is inserted which restricts
movement (Rocscience, 2008). Within the Hutchinson Joss Bay model, it was found
that the condition assigned to the end of the joint did not influence the subsequent

117
CHAPTER 5: Case study (chalk) slopes 2 and 3

SRF, however more plastic strain did result at the end of a closed end joint presumably
due to restricted movement.

Importantly mesh density has a significant influence on the SRF, with a finer mesh
leading to a smaller SRF, as depicted in Figure 5.3, presumably due to the ease of
strain development and localisation. The coarse model revealed a SRF of 1.15,
indicating that the model was above the limit of stability. With a manually-inserted finer
mesh the mass strength needed to be significantly higher, with the initial SRF between
0.6 and 0.8, consequently this could indicate that the mass  and c to be 62° and
44kPa respectively (if SRF taken to be 0.6). Therefore the mass properties that Hoek
and Bray (1981) report could be below the limit point of failure when there is a pre-
existing tension crack.

Figure 5.3: Mesh dependency and consequent different behaviour within the
2
Hutchinson Joss Bay Phase model, with a pre-existing tension crack.

Within a Phase2 model with no tension crack, fine mesh and extended boundaries,
shown in Figure 5.4a, an SRF of close to 1 was predicted (using the mass strength
properties listed in Table 5.5). Therefore the inclusion of a pre-existing tension crack
within a finely meshed model can have a significant influence on the back-analysed
rock mass strength. However, it must be noted that in none of the Phase2 SSR
simulations, was tensile strength included in the SSR process. It is suggested that the
118
CHAPTER 5: Case study (chalk) slopes 2 and 3

influence of this should be reviewed, as this may significantly affect the SRF. In
addition the finely meshed Phase2 model with no tension crack, clearly demonstrates
that the FEM/BEM in Phase2 can predict the locality of developing tension cracks (prior
to a reduction of the strength beyond a SRF of 1), as illustrated in Figure 5.4b.

Figure 5.4: Model extent (a) and


subsequent ability of Phase2 to predict
tension crack development (b) within
the Hutchinson Joss Bay model, with
no pre-existing tension crack.

Further investigation into the appropriate mesh size, is necessary if the results from this
study are to be used in any further analysis. Also it is recommended that the strength
applied to the tension crack should be reviewed. In particular the influence of the
stiffness applied to the tension crack was not studied; this may increase the FOS found
within the model with a pre-existing tension crack.

The modelling reported in this section has demonstrated that it is possible to model the
failure mechanism within the Hutchinson Joss Bay failure using Phase2. However in
order to use a more comprehensive numerical model, such as Phase2, more data is
required, which can be problematic. Phase2 has proved to give a further insight into
the mechanisms and stress conditions within the slope, in comparison to the simple
limit equilibrium methods described in the previous section. To improve reliability of a

119
CHAPTER 5: Case study (chalk) slopes 2 and 3

more complex model, it is suggested that a more detailed review is necessary, to clarify
the failure geometry, existence of the tension crack and the intact strength of the
respective Chalk formations.

5.2.4 Results from FEM/DEM (ELFEN)


By using ELFEN, the development of shear strain and consequent tension can be
simulated. Several models have demonstrated that a shear plane extends from the
notch within the base of the model. Ideally a shear plane would be inserted when the
shear strain results in plasticity, however as discussed in Section 2.2.4 ELFEN cannot
extend fractures which are formed from mode II behaviour. Instead, as shown by
Klerck et al. (2004), shear failure can be indicated using the current version of ELFEN,
by tensile fractures that open within a shear band. This has not occurred within the
models presented in this section due to softening parameters not being assigned,
which would otherwise come into effect post-yield and cause tensile fracturing. Instead
for this research, ELFEN has been used only to analyse the peak strength of the rock
mass.

In addition to the parameters outlined in the previous sections, the fracture energy (Gf)
is required within ELFEN. This can be calculated from the fracture toughness (KIC)
which is derived from the intact σt using the equation presented in Appendix A.

Consequently the KIC for chalk can be estimated to be 0.218MPa m , assuming an


intact σt of 1.5MPa as suggested by Lawrence (2006). As a note, Whittaker et al.
(1992) do not publish any values for the KIC of chalk, instead the closest lithology is

limestone, which ranges from 0.66 to 4.20MPa m for limestone. Therefore the KIC
calculated for chalk is plausible, being considerably weaker than limestone. From the
KIC, a Gf of 47.5 J/m2 can be derived, as illustrated with the following calculation:

K IC 2 (218 kPa m) 2 47524 kPa 2 / m


Gf = = = = 0.0475 kPa/m
E 1000000 kPa 1000000 kPa
[5.2]
= 47.5 Pa/m = 47.5 N/m = 47.5 N  m/m = 47.5 J/m
3 2 2

The constitutive model chosen to represent the mass strength is the Mohr-Coulomb
with tensile Rankine cut-off, as discussed in Appendix B. A model with a purely
Rankine criterion was simulated, resulting in no failure due to the fact that the mass-
controlled shearing cannot be considered. If a purely Mohr-Coulomb model was used,

120
CHAPTER 5: Case study (chalk) slopes 2 and 3

only shear failure can take place and tensile fracturing will not occur. The mixed mode
I-II failure mechanism, as discussed in Section 2.2.4, therefore demonstrates the
requirement of the coupled criterion, when using a fracture-based code such as
ELFEN.

Mesh dependency
Elmo (2006) reported on post-peak mesh-dependency within 3D uniaxial ELFEN
models, where a mesh density of 1mm was required to predict shear bands (5mm
meshed model had no shear bands). This section reports on a similar mesh-
dependency however concerning pre-peak behaviour. During the modelling of the
Hutchinson Joss Bay model 1, it was found that a particular mesh density was required
to allow the full development of shear-strain up to the tension crack. As illustrated in
Figure 5.5b a mesh of 0.4m (within the mesh spheres detailed in Appendix L) was too
coarse; instead, as shown in Figure 5.5c full development of the shear-strain only
occurred with a finer mesh of 0.3m (within the mesh spheres).

Also Figure 5.5d demonstrates that further development of strain within the tension
crack and consequent deformation, only occurred when the ‘tetrahedral element’
(option 37) control is assigned. There is no guidance within the literature supplied by
Rockfield (2008), however it has been suggested by Rockfield (2007, personal
communication) that option 37 is necessary for slope projects where plasticity is going
to occur. Option 37 controls the computation of the stresses within the model, allowing
an algorithm to be used which improves the stress distribution across elements,
compensating for the over-stiffness of triangular elements (Rockfield, 2007, personal
communication). It is recommended that there should be further research into the
influence option 37, as Rockfield (2007, personal communication) state that the
algorithm is still under development.

The material properties used in the ELFEN simulations of the Hutchinson Joss bay
models, were all based around those that were used in Phase2, which were listed in
Table 5.5. For the further mesh-based and runtime parameters that were used in each,
see Section ii of the model database.

121
CHAPTER 5: Case study (chalk) slopes 2 and 3

Figure 5.5: ELFEN Hutchinson Joss Bay model 1, illustrating the meshed regions (a); development of
shear strain when mesh density within the spheres = 0.4m and 0.3m (b) and (c) respectively; (d)
addition of ‘Option 37.’

Stress-path monitoring
During simulations, it was noted that there were different stresses at nodes in
comparison to elements. In particular a peak occurred in principal stresses, at the start
of each of the excavations, when the stress was monitored at nodes. Rockfield (2007,
personal communication), suggested that this is a limitation in the current post-
processor version of ELFEN. Consequently to avoid monitoring the nodal stresses
within each of the independent stages (and subsequently joining them together),
elemental stresses are monitored instead, as the stress paths within elements remains
unaffected by the excavation sequence. Note that points for element monitoring have
to be assigned prior to the simulation, unlike nodal monitoring. This can hinder the
detection of failure; as discussed later in this section stress can be locally different
between models depending on slope unloading mechanism, and other slight changes.
This is due to the sensitivity of the FEM/DEM (see Section 2.3) modelling scheme,
which can restrict comparisons between models where elemental stress path
monitoring is performed.

Primarily, simulations were conducted using Hutchinson Joss Bay model 2, with a pre-
existing tension crack using various tensile strengths. A relatively high tensile strength
(σt) of 1.5MPa, suggested by Lawrence (2006) as a viable intact σt, causes fracturing
instantaneously at 9s, as illustrated in Figure 5.6a. Note that failure occurs within this
model following the relaxation of the final excavation object (between 8 and 9s).
122
CHAPTER 5: Case study (chalk) slopes 2 and 3

When the Hutchinson Joss Bay model 2 was run with a lower, more likely mass σt of
13kPa, as derived in Section 5.2.3, the behaviour of the model is considerably more
ductile. Tensile failure occurs after the final excavation, as presented in Figure 5.6b;
also the fracturing is more gradual, occurring primarily at the base of the tension crack
due to the ductile deformation of the face, followed by further fracturing within the toe of
the slope.

Figure 5.6: ELFEN Hutchinson Joss Bay model 2, illustrating the dependency on tensile
strength. (a) (b) Mohr-Coulomb stress paths at the base of the tension crack, and screenshot of
final fractured state, when the σt is 1.5MPa and 0.013MPa respectively.

Importantly within all models, shear strain initiated within the toe following release of
the slope constraint, and subsequently progressed upwards. Within a model with no
tension crack (Hutchinson Joss Bay model 3), it was demonstrated that tensile failure
can result from the progressive development of shear strain. Consequently a tension
crack formed from 5.5s, (0.5s after the final release of the slope face, which finishes at
5s), as illustrated in Figure 5.7.

123
CHAPTER 5: Case study (chalk) slopes 2 and 3

Figure 5.7: Progressive development


of shear strain and subsequent
opening of a tension crack within
ELFEN model of Hutchinson Joss Bay
cliff section.

Although Figure 5.7 shows shear strain to develop from 5.1s, the precise time of failure
was investigated at two monitoring locations, Point A and B illustrated in
Figure 5.8. Figure 5.8a shows those elements that have proceeded past yield point
into plasticity at 5.5s when the tension crack starts to form. These form two shear
bands which are not detected on a shear strain plot. It was discovered, (following the
selection of the monitoring points), that point B lies between these two bands of plastic
elements. Consequently failure did not occur at point B within this particular model,
instead shear failure can only be tracked within point A, occurring from 4.7s, as
illustrated in Figure 5.8b. Tensile failure at point B was noted within a later model and
is presented within the following section (Figure 5.11).

Figure 5.8: ELFEN Hutchinson Joss Bay model 3, demonstrating the detection of failure at point A;
which is within the zone of failed elements (a) as shown on Mohr-Coulomb stress-path plot (b).

124
CHAPTER 5: Case study (chalk) slopes 2 and 3

Method of unloading
As discussed in Section 3.1.1 the release of the Hutchinson Joss Bay slope, by
relaxation of a structural-constraint, was not possible (failure could not occur) within
models where there is a tension crack. This is due to the meshing algorithm within the
version of ELFEN used (analysis 3.8.5), which inserted a ‘beam element’ into the
‘mei’ file (Rockfield, 2006, personal communication).

These models were simulated prior to the suggestion from Lobao (2007, personal
communication) that instead of the structural-constraint, an applied displacement load
of zero can act as the constraint during the constraint-release approach. Consequently
slope failure, within the models where there is a pre-existing tension crack, occurred
due to gravity loading or release of a single or several excavation object(s). Therefore
the influence of the staged excavation process has to be considered, as reported
below.

Different methods of unloading were investigated, to determine the influence on the


resulting stress state within the Hutchinson Joss Bay model where there is no pre-
existing tension crack. Four models were used (Hutchinson Joss Bay models 4 to 7),
each with an independent unloading sequence, as shown in Figure 5.9 (additional
detail for each model, is presented in the model database). All simulations used the
basic material properties listed in Table 5.5; other parameters and the specific mesh
detail for each model, is as presented in the model database.

Figure 5.9: Hutchinson Joss Bay unloading models.

Each model produced a slightly differently stressed model; this also resulted in zones
of elements that had reached plasticity, which are individual to each model.
Figure 5.10a presents the distribution of these zones at the time that the band of
elements reaches the top of the slope, within each model.

125
CHAPTER 5: Case study (chalk) slopes 2 and 3

In addition, as previously discussed, a shear plane cannot be inserted into the model
due to a limitation of the current version of ELFEN; therefore an ELFEN Hutchinson
Joss Bay model will only run until deformation within the resulting tension crack is too
much to model. It seems that this final stage that the model runs to is dependent on
the unloading method, as presented in Figure 5.10b. It is the vertically staged
excavation that shows the most advanced form of deformation, and is consequently
included as the video file “Hutchinson Joss Bay model 6” on the appended CD-rom.

Figure 5.10: Hutchinson Joss Bay models 4 to 7 (from left to right); (a) runtime, at which the
elements that have reached plasticity intersect the top of the slope. (b) Final fractured state
of each model.

Despite their similar final-stress results, the way in which each of the models reached
these stress states is different. The most important finding from stress monitoring, is
that failure does not occur at point A within the horizontal and vertical excavation
models, or B within the vertical and single excavation model. Failure therefore only
occurs at both monitoring points, within the model which failed due to purely gravity
loading (non-excavation model). Therefore it is recommended that stress monitoring
within future simulations should be set to be within many locations to attempt to capture
the failure process, regardless of the small stress variances between each of the
unloading techniques. Figure 5.11 presents a Mohr-Coulomb stress path for both
points A and B within the non-excavation model; as detailed in model database, gravity
loading is from 0 to 1s, with 1 to 2s being a period for the release of any potential
discontinuities, after which progressive failure occurs from toe of the slope.
126
CHAPTER 5: Case study (chalk) slopes 2 and 3

Figure 5.11: (a) (c) Plasticity within ELFEN Hutchinson Joss Bay model 4, demonstrating the primary
detection of compressive failure (b) at point A; followed by tensile failure at point B as shown in (d).

As is illustrated by the zones of plastic elements within Figure 5.10a, the stress
variances within models is localised. Globally stresses are similar between the models,
regardless of unloading method; therefore out of all of the unloading methods,
excavation in one stage is preferable as this provides a rapid suitable way of accurately
stressing and unloading a model, as discussed in Section 3.1. However, a single
excavation model is not as kinetically stable as a staged excavation model, which may
lead to a less-developed fractured profile.

Shear strength reduction


The single-excavation model (Hutchinson Joss Bay model 5), was used for the SSR
study within ELFEN, which suggested that the mass strength (c = 26.4kPa, and
 = 50°) presented by Hoek and Bray (1981), could be slightly below the limit point.
Consequently Mohr-Coulomb mass strength parameters had to be increased slightly,
(via a ‘shear strength increase factor,’ used as a denominator in same way as the SRF
is used in Equation 2.4), to find the limit point, as illustrated in Figure 5.12.

127
CHAPTER 5: Case study (chalk) slopes 2 and 3

Figure 5.12: Analysis of mass strength within an ELFEN Hutchinson Joss Bay
model 5, demonstrating a shear strength reduction factor of close to 1. Final
degree of plastic strain is plotted for each of the mass strengths.

5.2.5 Conclusions from Hutchinson Joss Bay Modelling


The analysis of the Hutchinson Joss Bay model using various limit equilibrium methods
has demonstrated that the slice-based computation within SLIDE, reveals a FOS which
is just greater than 1. The alternative discrete-block analytical methods, within
Plane_Failure and RocPlane, both returned a FOS that is closer to 1.1; this could be
due to the fact that simplified geometries had to be created. As a recommendation, the
Microsoft© Excel-based Jacob (Pine, 2006a) should be considered for future non-
circular analysis, as in Chapter 7, to provide further verification of the results from
SLIDE.

The more advanced numerical analysis methods (Phase2 and ELFEN) give a FOS of 1
and just less than 1 respectively, (in models without a pre-existing tension crack). In
addition both Phase2 and ELFEN modelling demonstrated that mesh density is a key
issue to the accurate analysis of a mixed mass-discontinuity failure mechanism. A
mass σt was derived using an empirical approach, which was shown within ELFEN to
enable a more realistic behaviour. Stress-path monitoring illustrated that tensile failure
within a model with a low σt followed a ductile response. In many models failure was
not permitted with a high σt; however within the Hutchinson Joss Bay model 2 (which
had a pre-existing tension crack), failure was achieved. However, analysis of stress,
demonstrated a high compressive stress where tensile failure should be occurring (as
illustrated in Figure 5.6a). As a consequence failure was instantaneous when σt was
high, indicating a rigid behaviour that is perhaps unrealistic for chalk.

128
CHAPTER 5: Case study (chalk) slopes 2 and 3

Further stress-path modelling was performed within ELFEN models with a lower and
more realistic tensile strength. Hutchinson Joss Bay models 3 to 7 all had no pre-
existing tension crack, designed to illustrate that a tension crack can form in the final
stages of failure. Stress-path monitoring showed early compressive failure within the
toe of the slope. Following this, the progressive development of shear strain upwards
led to tensile failure. This demonstrated the ability of ELFEN to simulate the
successive development of a mixed mode I/II failure mechanism, with an indication of
slope damage by tensile fracturing in the locality of the tension crack.

Different unloading methods were simulated, and stresses were monitored. Localised
differences within models were noted, however globally the stress and failure situation
within each model was the same. There could be an apparent link between kinetic
stability, (related to the method of unloading), and final fractured state to which ELFEN
will proceed. This could illustrate that a slow method of unloading (in this case staged
excavation), permitted further slope damage; however further work is needed to
confirm the observations made.

A SSR process within Phase2 and ELFEN proved that the Hutchinson Joss Bay model
has close to the limit of stability when there was no pre-existing tension crack. Due to
the logistical time-frame, the SSR method was applied to the model with a pre-existing
tension crack, only using Phase2. Results suggested that the influence of a pre-
existing tension crack was significant, leading to a much higher assumption for rock
mass strength. However further study into this is required, with consideration of the
tensile strength during the SSR process and the strength applied to the tension crack.

Although Hutchinson (1970 cited in Hoek and Bray, 1981) suggested failure was under
drained conditions, future investigation should perhaps detail the influence of
groundwater. Groundwater conditions are especially critical to the stability of chalk
slopes; Mortimore et al. (2004b) note a marked reduction in chalk strength, with
increased saturation; for instance they found that Brazilian strength can be up to a
quarter when the specimen is saturated. This could be due to the small-scale
structure/fabric that chalk can have, which upon saturation is mobilised resulting in
reduced intact strength. The micro-fabric within chalk was also suggested by
Mortimore et al. (2004b) to be the cause of the non-uniform behaviour during triaxial
tests. They found that conventional Coulomb theory, which states that failure should
take place in planes at 45° + /2, could not be applied to chalk; instead failure planes
frequently developed along pre-existing fabrics such as wisps of marl, a vein fabric or a
fossil.

129
CHAPTER 5: Case study (chalk) slopes 2 and 3

With the current analyses that are available, the simulation of groundwater within a
mass controlled instability such as the Hutchinson Joss Bay model is restricted to limit
equilibrium and FEM approaches. In order to consider groundwater within a chalk cliff,
using the current version of ELFEN, a more discontinuity-based slope failure has to be
analysed. This is due to the way in which ELFEN simulates the influence of pore-
pressure, as outlined in Section 2.4.2. Subsequently the development and analysis of
a discontinuity-controlled chalk model, is outlined in the following section.

5.3 Case Study 3 – Theoretical step-path failure through chalk


The following section details the development of a theoretical step-path model of a
chalk cliff, with the aim of testing the current effective stress module within ELFEN. As
discussed in Section 2.4.2 a discontinuity controlled slope is required, in this case a
step-path geometry was developed where kinematic release is possible. Section 5.3.1
describes the time consuming process that was required. This involved the calibration
of slope geometry with rock-bridge and discontinuity strength, such that under drained
conditions the slope is stable, however following saturation instability subsequently
arises.

5.3.1 Model Geometries


It is a difficult task to create a fractured (step-path) slope that is stable in the dry, but
fails due to a degree of saturation; such a model has to be near the limit of stability. In
addition the achievement of full kinematic release as a result of step-path failure is also
problematic. Four different models were created by trial and error development; all
simulations used the same basic model properties outlined in Table 5.6.

Table 5.6: Basic material and discontinuity strength properties used in chalk step-path models.

Property Value
Young’s Modulus (GPa) 1
Main Properties: Poisson’s Ratio 0.24
Density (kg/m3) 1700
Cohesion (kPa) 26
Mohr-Coulomb and
Friction angle (°) 50
Rotating crack properties
Dilation angle (°) 5
(representing mass
Tensile Strength (kPa) 13
strength):
Fracture Energy (J/m2) 47.5
Cohesion (kPa) 0
Discontinuity-based Friction (°) 20
strength properties: Normal Penalty (kPa/m) 200
Tangential penalty (kPa/m) 200

130
CHAPTER 5: Case study (chalk) slopes 2 and 3

In addition to the properties listed in Table 5.6, groundwater parameters had to be


incorporated into the ELFEN model. These are presented in the model database and
also in Appendix K, where some inconsistencies are discussed (unfortunately
discovered after the modelling).

Model 1, presented in Figure 5.13a, revealed instability in drained conditions.


Appendix I describes development of the model, addressing restrictions in kinematic
release and realistic fracturing. This led to the simulation of a successful step-path
fracture and kinematic release, (video file “Chalk step-path model 1” on the appended
CD-rom). Although model 1 demonstrated failure, it was under unsaturated conditions;
in which case for the specified rock-bridge and discontinuity strength shown in
Table 5.6, the inclination of the slope was too steep to achieve stability. Consequently
model 2 (Figure 5.13b) was created, which showed the opposite situation, where
despite full saturation, failure would not occur, most likely due to an insufficient slope
angle.

Prior to the re-evaluation of discontinuity and rock-bridge strength, model 3 was


created, by increasing the dip of the slope face. This simulation exhibited failure under
dry conditions, with incomplete fracturing through some of the rock-bridges, as
illustrated in Figure 5.13c. This meant that with the particular set of conditions used in
model 3, kinematic release was not possible, possibly due to a limitation within the
current version of ELFEN (inability to fracture under mixed mode I-II failure).

However, as Figure 5.13c shows, model 3 predicts a tension crack forming in front of
the step-path fractures, in a similar location to the tension crack noted within the
ELFEN models of the Hutchinson Joss Bay cliff section, presented in the previous
section. This could indicate that the pathway for step-path failure needs to be steeper
(expected for chalk, discussed in Section 2.2.5), as within the other three models.

Consequently the slope angle of model 1 was decreased and step-path fractures re-
aligned slightly, to ensure a slightly shallower failure surface than in model 1. This
created model 4, presented in Figure 5.13d. This model provided the opportunity to
test the effective stress module within ELFEN using a fracturing model; the results from
which are presented in the following section.

However, the degree of kinematic release that model 1 exhibited was not achieved
within model 4, as shown in the video file “Chalk step-path model 4” on the appended
CD-rom. Therefore it could be suggested that a full failed state was not attained even

131
CHAPTER 5: Case study (chalk) slopes 2 and 3

at 100% saturation and further calibration is required to achieve kinematic release.


Despite this, a brief sensitivity study was conducted using model 4, as outlined in the
following section.

Figure 5.13: ELFEN models of step-path failure within an example 15m high chalk cliff: (a) model 1 was
restrictive (see Appendix I); (b) no fracture across rock-bridges occurred in model 2; (c) kinematic
release could not be achieved in model 3; (d) Final model which revealed successful results

5.3.2 Results from phreatic surface modelling


It was found that failure occurred within model 4, when the phreatic surface increased
from a height of 80% saturation within the slope, to 100%, as shown within Figures
5.14a and 5.14b respectively.

Figure 5.14: ELFEN model step-path model 4, illustrating instability and consequent step-
path failure as a result of rise in pore pressure (displayed in N/m2, with the phreatic surface
superimposed).

The sensitivity of the slope failure, to the frictional strength () applied to discontinuities,
was briefly studied. As discussed above, failure occurred when the slope became
saturated with greater than 80% groundwater, with the discontinuity  at 45°. When the
discontinuity  is decreased to 35°, instability occurs when the slope is more than 50%
132
CHAPTER 5: Case study (chalk) slopes 2 and 3

saturated. However as illustrated in Figure 5.15, there was a limited amount of rock-
bridge failure.

Figure 5.15: ELFEN step-path


model 4 illustrating the final fractured
state with low discontinuity frictional
strength and 50% saturation.

Further simulations could be conducted to provide a range of discontinuity strengths


and phreatic surfaces with which the theoretical slope becomes unstable; this would
require further manual plotting of the phreatic surface co-ordinates. Instead, the
dependency of slope stability on mass tensile strength was studied briefly, to provide
an insight into the fracturing process through the rock-bridges. The data from this
study is not presented due to an unrealistic brittle response when σt was increased
from 13kPa to 50kPa.

In conclusion the step-path model proved that the fracturing across rock-bridges is
especially dependent on the frictional strength applied to step-path discontinuities, and
the rock mass strength. In addition it was demonstrated that either a high σt and/or a
high discontinuity  can make a model more unstable. It is recommended that further
work is necessary on a stronger model, to fully assess the degree of kinetic energy
associated with loading and release of a fractured slope, as discussed previously in
Section 3.2.

5.4 Summary
A range of methods were used to analyse the stability of the Hutchinson Joss Bay
failure, different results were obtained from slice-based and discrete limit equilibrium
analyses. The slice-based (SLIDE) model predicted instability close to limit point (with
an average FOS of 1.05), whereas stability was predicted within two discrete-based
analytical methods (RocPlane and Plane_failure), with a FOS of 1.1. It is suggested
that further work is conducted to provide an accurate representation of the Hutchinson
Joss Bay failure using limit equilibrium techniques.

A more advanced analysis of the model, using the FEM/BEM approach (Phase2) and a
comprehensive FEM/DEM fracture-based approach (ELFEN), demonstrated similar

133
CHAPTER 5: Case study (chalk) slopes 2 and 3

results. Rock mass strength was analysed using both analyses methods, via a SSR
approach (only within Phase2 is the SSR automated). This showed rock mass strength
was near limit point within a model with no pre-existing tension crack. A SSR
2
simulation was performed in Phase using a model with a pre-existing tension crack,
showing that this could significantly influence the back-analysed rock mass strength.
Further analysis of this particular model is required, using the SSR approach to provide
confidence in rock mass strength.

Importantly this investigation has shown that elements of the Hutchinson Joss Bay
failure can be simulated using a limit equilibrium method, with further detail provided by
the subsequent more sophisticated levels of analysis. Groundwater was not considered
within the Hutchinson Joss Bay model, on account that failure was under dry conditions
(Hutchinson, 1970 cited in Hoek and Bray, 1981). However from review of literature, it
is clear that moisture can have a significant influence on the stability of chalk cliffs. In
general groundwater can have an influence on intact strength, due to moisture affecting
cohesive and frictional components of intact chalk strength. More critically there is also
the reduction of shear strength as result of groundwater pressure reducing the normal
effective stress.

As discussed in Section 2.4.2, the present version of ELFEN can only simulate the
influence of pore pressure on the shear stress within discontinuities. Therefore to
investigate the performance of ELFEN further, a discontinuity-controlled (step-path)
chalk cliff model had to be developed in preference to the mass-controlled Hutchinson
Joss Bay model.

Eventually a suitable model was developed, demonstrating that as a direct result of a


rise in the phreatic surface, a degree of step-path failure can be simulated. This proves
the ability of ELFEN, to simulate tensile fracture extension and consequent slope
instability as a result of a decrease in shear strength, relative to pore pressures within a
slope model.

Finally it is suggested that to fully capture the critical condition of groundwater within
chalk, discussed in Section 5.2.5, further small-scale characterisation of chalk is
required. Subsequently this could be used to perform biaxial tests within ELFEN, in a
similar manner to that presented in Chapter 8; however for chalk, rock mass strength
for different moisture conditions would be of value. In addition, further testing of the
groundwater module within ELFEN is necessary; this is performed within the following
Chapter.

134
CHAPTER 6: Case study slope 4 (Delabole)

CHAPTER 6: Case study 4 – modelling of a dominantly


discontinuity-controlled failure at inter-bench scale

6.1 Introduction
This Chapter outlines a back-analysis of a slope failure at Delabole slate quarry,
Cornwall. Three different numerical modelling methods, (RocPlane, SLIDE and
ELFEN), were used, with the main emphasis of the research on ELFEN, using models
to achieve aims 1 and 2 outlined in Section 1.2. Section iv of the model database
contains detail on the model parameters used in ELFEN, and Appendix L provides a
record of the boundaries and simulation files.

The failure that occurred in 1967 at Delabole, was dominantly discontinuity-controlled.


Coggan et al. (2007) suggest that there have been several failure mechanisms
proposed; one of the more likely mechanisms is illustrated by Coggan and Pine (1996),
who used UDEC to identify the driving mechanism to be a “chisel effect,” with the
resulting interaction between four major blocks, as discussed in Section 2.3.6. The
rotation and translation that occurred during the chisel effect, was driven by the upper
two blocks slipping down, which caused the block above the lower wedge to move up
and out from the surface and rotate.

The pre-failure detail suggested by Clover (1978, cited in Coggan and Pine, 1996)
clearly shows the geometry of the slope face and discontinuities involved, as illustrated
in Figure 6.1a. The pre-failure geometry can be simplified to represent the block
interaction that Coggan and Pine (1996) identify from UDEC modelling. Subsequently
it could be suggested that the mechanism is a form of active-passive wedge, with an
underlying element of toppling, as illustrated in Figure 6.1b.

Coggan and Pine (1996) review the locally named discontinuity sets, which have been
identified within the quarry; two sets of persistent faults that they present, strike sub-
parallel to the slope face. As illustrated in Figure 6.1 one of these sets of faults, the
claylodes, dip steeply into the slope. The other set of normal faults that occur within
the plane of the section through the west face, is the ‘shortahs’ which dip sub-parallel
to the face. The complete discontinuity forming the rear failure within the Delabole
model presented in Figure 6.1b is referred to as a shortah during this research; in
reality the release plane is more uneven as illustrated in Figure 6.1a, with perhaps
failure dominantly occurring along the shortahs but also through the mass.
135
CHAPTER 6: Case study slope 4 (Delabole)

Consequently, although the failure at Delabole was dominantly discontinuity-controlled,


this Chapter outlines an investigation into mass-controlled failure via the location of
rock-bridges within the base-plane, and on the shortah and claylodes. As shown with
Section 6.2, the ELFEN model which presented the most realistic failure mechanism,
included extended claylodes (as illustrated in Figure 6.1) to provide an aspect of
toppling. Also, within this particular model rock-bridges were located within the base-
plane and the shortah, to allow progressive failure and subsequent block interaction.

Figure 6.1: 1967 failure at Delabole slate quarry, Cornwall; (a) east-west section from Clover
(1978, cited in Coggan and Pine, 1996); (b) conceptual model illustrating the driving processes.

Following the UDEC simulations reported in Coggan and Pine (1996), there was a
move towards simulating the Delabole failure using a fracture-based code, to provide
more insight on the mechanism and processes controlling the failure. Coggan et al.
(2003) report on one of the first ELFEN models of the Delabole failure; the research of
this thesis followed on from their investigations with development of the ELFEN model.
Subsequently early models consisted of multiple surfaces separated by discrete
fractures, as illustrated in Figure 6.2.

This model construction allows the strength of the main slope to be increased,
constraining fracture to mainly take place within the failure blocks. This prevents the
136
CHAPTER 6: Case study slope 4 (Delabole)

otherwise unrealistic extensive fracturing behind the failure blocks, due to the dynamic-
response of the model (see Section 3.1), which was loaded rapidly and without a
constraint upon the slope face. Importantly the failure blocks were given weaker
properties to simulate the interaction of discrete failure blocks and their subsequent
comminution, which was originally reported by Coggan et al. (2003).

Figure 6.2: Original ELFEN model of the Delabole 1967 failure used by Coggan
et al. (2003).

However, the multiple surface modelling approach restricts the simulation of Delabole
to a purely discontinuity controlled failure mechanism. To provide a more realistic
representation of the processes, a ‘single-surface’ model (where fractures are
embedded), was developed. This allowed the development of fracture through the
mass, fulfilling the primary aim of the Delabole modelling during this research which
was to investigate the extent of fracture surfaces, especially the basal fracture. The
following section discusses the development of a single surface model.

6.2 Simulation of failure mechanism


Ten different ELFEN models are presented throughout this section; all of these used
the basic model parameters listed in Table 6.1. Section iv of the model database
outlines the detail regarding additional parameters that were used in each model.

137
CHAPTER 6: Case study slope 4 (Delabole)

Table 6.1: Summary of the most important properties used for the mass and discontinuity
strength in Delabole models (for further parameters see Section iv of the model database).

Parameter Value
Young’s Modulus (GPa) 20
Main Poisson's ratio 0.25
properties
Density (kg/m3) 2500
Mass
Mohr- Cohesion (MPa) 2
strength
properties Coulomb Friction Angle (°) 46
and Dilation angle (°) 5
Rotating Tensile strength (MPa) 0.9*
crack Fracture energy (J/m2) 20
Friction (°) 31
Cohesion (MPa) 0.01**
Discontinuity Discrete
strength global Varied between 0.1 and
Normal Penalty (GPa/m)
properties properties 1
Varied between 0.01
Tangential Penalty (GPa/m)
and 0.1

* The tensile strength for the mass was reduced to 0.756MPa in Delabole
model 2; and increased to 2MPa from model 4 onwards.
** From Delabole model 3 and onwards, the discontinuity cohesive strength
was reduced to zero.

The early single-surface Delabole models followed the geometry illustrated in Figure
6.3a, however the particular geometry (model 1) proved very sensitive, using the
properties listed in Table 6.1. The particular model was kinetically unstable, with
fracturing occurring as soon as the material retaining the face is excavated. The model
would then stabilise, with no further propagation even when the excavation object
restraining the face is fully relaxed (excavated).

In particular, kinetic instability within model 1 was expressed by failure occurring during
the loading stage, as illustrated in Figure 6.3a. Although the fracture in model 1 is
dynamic-related, the fracture pattern is relatively realistic (Figure 6.3b), until the stage
where excavation object is fully relaxed. It is at this stage when kinetic instability is
expressed by extensive fracturing, as illustrated in Figure 6.3c.

To improve the kinetic situation within the simulation, model 2 was created which had
extended boundaries, (see Appendix L). This did provide a dynamically more
appropriate model, without failure occurring during the loading stage; however the
model revealed a final state of very limited fracture, as presented in Figure 6.3d.
Therefore with the parameters used in model 1 and 2, (see model database), it was
assumed that complete failure through the rock-bridges could not be attained.

138
CHAPTER 6: Case study slope 4 (Delabole)

Figure 6.3: (a) (b) (c) Fracturing during respective modelling stages of model 1. (d) Shows model 2
where simulation shows kinetic stability, although fracture does not extend through the rock-bridge.

To promote instability, model 3 was developed with extended claylodes to allow


toppling as illustrated in Figure 6.4a (in addition normal penalty applied to all
discontinuities was reduced by a factor of ten). With a rock-bridge between the basal
fracture and shortah, kinematic release was still not achieved, despite the rock-bridge
failure illustrated in Figure 6.4b. Note that the failure state illustrated in Figure 6.4b
was only achievable via a model where failure occurred due to gravity loading (with no
excavation object). In the same model which failed due to an excavation process, the
fracture through rock bridges was intense causing the simulation to cease; clearly the
loading regime requires review within the excavation model.

Figure 6.4: (a) Geometry of


model with extended
claylodes (Delabole model 3),
to promote toppling. However
fracturing through rock-
bridges appears intensive (b)
with no kinematic release.

Following the unsuccessful attempts to get the basal fracture to extend, a step-path
geometry was created, firstly within a model where there were incomplete claylodes as
illustrated in Figure 6.5a. In addition the strength criterion was altered, so that
Delabole models 4 to 10 only used a purely Rankine with rotating crack formulation, as
opposed to the Mohr-Coulomb and Rankine coupled criterion (see Appendix B). It was
also found that the tensile strength for the mass, had to be increased to 2MPa.

139
CHAPTER 6: Case study slope 4 (Delabole)

Fracture over the step-path was achieved within an excavation model, however no
further fracturing across the failure blocks occurred. On re-examination of project files,
the limited kinematic release and subsequent incomplete failure mechanisms noted in
the above (excavation) models may be due to the basal fractures not extending
through the slope face. As discussed in Section 3.1.1 and in Appendix F, to achieve
kinematic release, embedded discontinuities have to extend through the slope face.
Within the Delabole models this appears to be case, but only within models where
there is an excavation object retaining the face. Consequently it is suggested that
some of the above models are re-run with extended basal fractures. Also at the time of
modelling, the constraint-release method had not been fully developed; therefore this
remains a technique that has not been used in these early Delabole simulations and
should perhaps be an area of recommendation for further work.

Despite discontinuities not extending through the slope face, it was found that
kinematic release could be achieved within step-path models where there is no
excavation object. However, primarily many models were kinetically unstable, leading
to numerous simulations where there was extensive dynamic related fracture;
consequently the failure mechanism could only be progressed to the state illustrated in
Figure 6.5b. When a more dynamically-stable model was developed (by loading over
4s as opposed to 1s), it appears that the progression of the model was heavily
dependent on the mesh density at which fracturing was permitted to occur. A model
where fracturing was permitted within a coarse meshed model, permitted a less
realistic post-failure response than a model with a fine mesh density, as illustrated by
comparing Figures 6.5c and 6.5d respectively.

Figure 6.5: (a) Step-path geometry, and final state of fracture achieved in model 4 (b). (c) Model 5, which
is more kinetically stable, but presents a blocky-type failure; whereas a slightly different element control
and finer mesh in model 6 (see model database), causes an alternative style of failure (d).

140
CHAPTER 6: Case study slope 4 (Delabole)

Importantly within all of the models presented in Figure 6.5, the block interaction,
suggested in Section 6.1, was not simulated. It appears that not only do step-path
fractures encourage kinematic release, but also the claylodes need to be extended in
order for the realisation of the toppling, translation and rotation behaviour. In addition,
several other issues had to be addressed during model development. Initially a more
realistic block interaction was exhibited by model 7, with rotation of the lower active
wedge, as illustrated in Figure 6.6b; however on further simulations the chiselling block
appears fixed. This was remedied by the implementation of option 166 (see
Appendix A) within model 8; also shortahs were extended although this does not
appear to have a significant influence but perhaps needs a more detailed study.

Despite the chisel effect within model 8, rotation was not observed; this may be due to
the degree of penetration. Subsequently model 9 was developed, where penetration
was prevented. It is difficult to be sure as to which parameter caused this prevention of
penetration; as the model database shows, model 9 had an increased field and zone
(see Appendix A for explanation of field and zone parameters).

Importantly rotation occurs in the top active wedge due to the progressive failure of the
rock-bridge between the shortah and base-plane, as illustrated in Figure 6.6d.
Removal of the rock-bridge (between the shortah and base plane) within the same
model, permits rotation within the lower active wedge as presented in Figure 6.6e,
causing the translation of the passive wedge.

Figure 6.6: (a) Initial step-path geometry when claylodes are extended; (b) results from model 7. (c) Model
8, where penetration prevents rotation. (d) Model 9 where different rotations occur when the rock-bridge is
finally removed (e).

141
CHAPTER 6: Case study slope 4 (Delabole)

As a final simulation of the failure mechanism, model 10 was created where failure
occurs through the mass, via the introduction of a rock-bridge within the shortah, as
presented in Figure 6.7a. As the simulation progressed, failure and subsequent block
interaction can be modelled, as presented from Figures 6.7b to 6.7d (a video file of
“Delabole model 10” is included on the appended CD-rom).

Figure 6.7: Simulation of progressive failure through the mass due to a chisel effect, resulting
in rotation and translation, achieved using Delabole model 10.

To implement groundwater within the code, one has to start at a relatively simple model
and subsequently develop its complexity. The following section details the results from
three additional models of Delabole, where a phreatic surface has been simulated.

6.3 Introduction of a phreatic surface and implication for


stability
To introduce groundwater into a dynamic fracture-based code, one has to follow an
appropriate loading and release procedure (see Section 3.1), so that instead of failure
due to gravity, the trigger process is a rise in pore pressures. Consequently the
conditions applied to the discontinuities that control the failure, need to be reviewed so
that a state close to FOS = 1, (see Section 2.2.2), can be simulated.

To consider the discontinuity strength, the Delabole failure was firstly simulated as a
planar situation, as introduced in Section 6.3.1. Following the calibration of numerical
parameters using the planar failure model, a rock-bridge is re-introduced within two
models detailed in Section 6.3.2. The first of these models is a relatively simple
situation with no additional discontinuities; importantly conditions are included so that
the rock-bridge fails as a result of a rise in pore pressure. Finally for a brief
consideration of an extended fracture network, a more complex model was simulated.

142
CHAPTER 6: Case study slope 4 (Delabole)

6.3.1 Delabole planar failure modelling


Several numerical methods can be used to include the effect of discontinuity strength.
The simplest approach is to use a limit-equilibrium method as introduced in Section
2.3.3; firstly a non-circular analysis in SLIDE (Rocscience, 2008) showed that a  of at
least 55° is required on the discontinuities if the model is to be stable during
unsaturated conditions (where there is no additional cohesive strength). In addition to
this slice-based approach, a discrete limit equilibrium method was used, RocPlane
(Rocscience, 2008). This predicted that the critical discontinuity  was lower (45°).
However, on review of the model it could be suggested that the discontinuity strength
assigned within RocPlane, is only applied to the base-plane because the shortah is
considered as a tension crack, as illustrated in Figure 6.8.

Figure 6.8: Geometry when simulating


the Delabole failure as planar
mechanism. Limit state solutions were
derived using RocPlane (Rocscience,
2008), which considers the rear-release
surface as a tension crack, with no
particular strength assigned.

The limit state solutions predicted by RocPlane were used to calibrate the behaviour
within the ELFEN model. However at the time, the potential lack of shear strength
within the shortah in the RocPlane analysis, was not considered; consequently within
the ELFEN model discontinuity strength was applied to both the base-plane and the
shortah. This possible disparity may account for some of the poor comparisons noted
in the following part of this section.

Firstly, as discussed throughout Chapter 4, the penalties assigned to the discontinuities


(see Appendix A) can significantly influence shear displacement. It was discovered
that there was no displacement in a wet model even though there was a low  (45˚)
applied to the discontinuities and phreatic surface was at a level where the shortah was
100% full, when the penalties were set at 200MPa/m and 20MPa/m (Pn and Pt
respectively). Consequently penalties were reduced to 20MPa/m and
2MPa/m, and displacement occurred. If no strength was applied to the shortah, then a
more unstable state would be predicted, in which case the penalties may be higher.
Higher penalties used during the calibration of damping, could potentially improve the
143
CHAPTER 6: Case study slope 4 (Delabole)

subsequent comparisons, between RocPlane and ELFEN, detailed later within Table
6.2. Therefore it is recommended that further research is conducted on the penalties
assigned to the discontinuities within the Delabole model.

During the calibration of damping within ELFEN, an under-damped situation was


discovered when using 5% point damping, as illustrated within Figure 6.9. This was
expressed by small-scale fluctuations in vertical displacement of the failure block when
the discontinuity strength defined a FOS of greater than 1; (in this case discontinuity
was purely frictional with a  of 60°, which in RocPlane equates to FOS of 1.8). This
fluctuation is likely to be related to the dynamic-response of the model, suggesting that
damping should be greater. As Figure 6.9 shows, there was significantly less
fluctuation in displacement within the same model with 10% damping leading to earlier
equilibrium.

Figure 6.9: Fluctuation dynamic-type behaviour indicated through the vertical


displacement after gravity loading, within an ELFEN Delabole planar failure
model where the FOS >1.

However it was found that 10% point damping inhibited full kinematic release when the
FOS was known to be lower than 1, as shown in Figure 6.10 (in this case discontinuity
was purely frictional with a  of 45° and shortah was 50% saturated, which in RocPlane
equates to FOS of 0.5). With 10% damping a degree of failure occurred but the failure
block did not fully release, suggesting the model was now over-damped because full
failure was achieved in the same model with 5% point damping.

Because both under-damped and over-damped situations were found to occur within
the ELFEN Delabole planar failure model, the staged approach had to be reviewed to
attempt to modify the dynamic behaviour presented within Figure 6.9, without

144
CHAPTER 6: Case study slope 4 (Delabole)

increasing the damping. The following section details how this was achieved, resulting
in a more kinetically stable model, which can be used for an accurate analysis.

Figure 6.10: Over-damped situation, when 10% point damping is used within
an ELFEN Delabole plane failure model, indicated by vertical displacement of
failure block when FOS <1.

Review of loading and release procedure


To decrease dynamic-related behaviour within a fracture-based model, without
introducing a higher degree of damping, one can either:

1. Increase the time over which gravity loading is applied; or,


2. Increase the time over which the constraint on the slope face is released.

Both of these methods lengthen the periods within which damping should diminish the
energy levels that cause dynamic effects.

Within the planar failure model that returned the results presented in Figures 6.9
and 6.10, gravity loading was applied via a drop-load curve and the slope face was
released over a period of 1s. A simulation was run where instead, loading followed a
ramped-curve but this made no difference to the energy levels within the model.

The second method presented above was more successful, decreasing the dynamic-
related behaviour. In particular when the release of the slope face is increased to
occur over 2s, a state of equilibrium in the kinetic energy is not achieved until
approximately 7s, with still a considerable peak in kinetic energy. Whereas if the slope
face is released over an even longer period of 4s, a state of equilibrium is reached
much sooner (by 5s), as illustrated by Figure 6.11. Importantly it appears that within
145
CHAPTER 6: Case study slope 4 (Delabole)

both models, kinetic energy peaks 1s prior to full release of the slope. The largest
peak occurs when the constraint on the slope face is released over 2s; this poses a
greater potential of dynamic-related behaviour. As discussed in Section 3.1.1, dynamic
responses within a model can lead to artificial failure.

Figure 6.11: Kinetic energy associated with releasing a constrained ELFEN


Delabole slope model where there is a discrete planar failure block
(discontinuity  = 60°).

From identifying a suitable degree of damping and duration of constraint release,


required within the Delabole model, the plane failure situation can be simulated with
more confidence using Delabole model 11. Material and discontinuity-related
parameters were as listed in Table 6.1, although in this case the penalties were ten
times lower, due to the results discussed in Section 6.3.1 (also discontinuity frictional
strength was applied, as outlined in Table 6.2). In addition, groundwater parameters
are presented in the model database and also in Appendix K, where some
inconsistencies are noted.

Simulation of groundwater within this model gave the results presented within Table
6.2. The FOS predicted by RocPlane (Rocscience, 2008), is also presented to allow
comparison.

As shown within Table 6.2, when a phreatic surface is applied to a model with purely
frictional discontinuity strength, the ELFEN model does not conform to the limit state
solutions from RocPlane; instead the FOS in RocPlane has to be between 0.5 and 0.6
for a FOS of 1 to be reached within ELFEN. This is likely to be due to RocPlane only
considering the frictional strength on the basal plane and not on the shortah, as

146
CHAPTER 6: Case study slope 4 (Delabole)

indicated earlier in Figure 6.8. Therefore, as discussed earlier, it is recommended that


numerical parameters and loading/release procedure are reviewed within an ELFEN
model with no strength applied to the shortah.

As well as potentially different discontinuity strength implied by RocPlane, it could be


suggested that there is a higher degree of saturation within the ELFEN models which
would alter the distribution of forces along the discontinuity in relation to pore pressure.
The simplified phreatic surface used within ELFEN needs to be more curved (requiring
a finer mesh density); this was also reported within Section 4.5.2 during planar failure
modelling. However it is likely that this would decrease the pore pressures acting
within the discontinuity and would subsequently increase stability. Therefore this would
contribute to an even more conservative behaviour within ELFEN, when compared to
FOS predicted by RocPlane. Consequently it is suggested that primarily discontinuity
strength applied to the shortah is reviewed.

Limiting conditions on
Vertical
discontinuities defining FOS Consequent
displacement of
discrete failure block predicted by FOS assumed

plane failure block
% water filling RocPlane in ELFEN**
at 10 s (m)
the shortah*
45° 0 (dry) 1.03 1.1 (stable) >1
43° 0 (dry) 0.97 8.5 (ongoing) <1
50° 30 0.90 0.62 (stable)*** >1
48° 30 0.83 0.92 (stable)*** >1
45° 30 0.75 1.7 (stable)*** >1
45° 40 0.62 2.1 (stable)*** >1
45° 50 0.48 2.2 (ongoing) <1

*Pore pressure within RocPlane is assigned following triangular distribution, with peak
pressure at the base of the shortah.

**The FOS within ELFEN is assumed from the degree of ongoing displacement noted;
as discussed in Sections 3.1.1 and 4.5, vertical displacement inevitably occurs in all
simulations when using a dynamic-based code. Where this displacement is minimal

FOS 1.
and a stable state is reached, it has been assumed that ELFEN has predicted a

***Displacement within saturated models appears to occur in two stages, one


associated with the release of the slope face, and the other related to the rise in pore
pressure.

The video file “Delabole model 11” included on the appended CD-rom, shows an
animation of the planar failure within ELFEN. This clearly demonstrates that there is a
primary displacement, which occurs as a result of the release of a constraint on the
slope face. The final trigger for the failure, and subsequent kinematic release, is a rise
147
CHAPTER 6: Case study slope 4 (Delabole)

in groundwater, in this case to fill the shortah by 50% (discontinuity frictional strength is
45°), as listed in Table 6.2.

Despite the poor comparison (shown in Table 6.2), between the limit state solution in
RocPlane and the behaviour in ELFEN, a fracturing model was created. As discussed
in the following section, a model was developed where kinematic release was achieved
within a fracturing model, as a result of rock-bridge failure, which was triggered by a
rise in pore pressure.

6.3.2 Delabole modelling with fracture due to groundwater


Model 12 was created, using the properties presented in Table 6.3; (for further
parameters see Section iv of the model database). Importantly a rock-bridge was
introduced into the base-plane, as shown later within Figure 6.12a.

Table 6.3: Summary of the most important properties used for the mass and discontinuity strength
in Delabole model 12.

Parameter Value
Young’s Modulus (GPa) 20
Main Poisson's ratio 0.25
properties
Density (kg/m3) 2500
Mass
strength Mohr- Cohesion (MPa) 2
properties Coulomb Friction Angle (°) 45
and Dilation angle (°) 5
Rotating Tensile strength (MPa) 1.5
crack Fracture energy (J/m2) 20
Friction (°) Varied between 36 and 38
Discontinuity Discrete
Cohesion (MPa) 1
strength global
properties properties Normal Penalty (GPa/m) 30*
Tangential Penalty (GPa/m) 10*

* It was found (as reported in the following section), that in order to get stability when
loading penalties had to be high, these were subsequently reduced to 0.02 and
0.002 GPa/m to allow unhindered sliding. Also for the same reason, the frictional
strength assigned to the base plane and the shortah was reduced to 36° after failure
had occurred.

The influence of the rock-bridge on kinetic energy effects was considerable; the staged
procedure developed within the previous section had to be reviewed to provide a model
where fracture within the rock-bridge does not occur, during loading and release. The
sequence developed was presented previously in Figure 3.1, as an example of staged
approach where a rise in groundwater is the trigger mechanism for failure. Similarly to
the sequence in model 11, the constraint upon the slope face was relaxed over a
period of 4s.

148
CHAPTER 6: Case study slope 4 (Delabole)

However in contrast to model 11, several other aspects of model were different due to
the sensitivity of the rock-bridge to dynamic effects. Kinetic energy induced by the
loading and rise in the phreatic surface, caused dynamic-related fracturing across the
rock-bridge prior to saturation of the discontinuities. Brief experimentation with the
duration of loading proved unsuccessful; instead it was found that certain modifications,
outlined below, helped to prevent this premature rock-bridge failure.

Modifications to step-path model to prevent dynamic-related failure of rock-


bridge
Fracturing across the rock-bridge during the loading stage proved difficult to prevent. It
was found that the implementation of Rayleigh damping (velocity and stiffness
proportional), as opposed to point damping (purely velocity proportional - see Appendix
A), provided an appropriate constraint on kinetic energy during the loading stage,
preventing the dynamic-related failure of the rock-bridge. This was only the case with a
ramped load-curve; if loaded instantaneously via a drop-load curve, then the kinetic
energy is too high resulting in the dynamic responses illustrated in Figure 6.12b.

Figure 6.12: (a) Stable situation when a ramped load curve is used within simulation of Delabole
model 12; (b) dynamic-related failure when the same model is instantaneously loaded.

It was subsequently discovered that although Rayleigh damping proved advantageous,


it can also act to restrict kinematic release by:

1. Inhibiting the failure of the rock-bridge when the phreatic surface rises
2. Significantly slowing the displacement of sliding blocks

Therefore Rayleigh damping should be replaced with point damping as soon as


possible in the staged procedure, following the loading stage. As discussed in Section
3.1.1, a settlement period should be included after the release of the discontinuities; it
is during this stage that point damping replaces the Rayleigh damping.

149
CHAPTER 6: Case study slope 4 (Delabole)

In addition, dynamic-related failure of the rock-bridge occurred in model 12 as a result


of the kinetic energy associated with the rise in the phreatic surface, even though the
discontinuities were not saturated, lying above the water table. A method of preventing
this dynamic-related rock-bridge failure is to increase the penalties assigned to the
base-plane by a factor of ten. However, as with Rayleigh damping, this modification
can inhibit kinematic release. Throughout Chapter 4 it was reported that high penalties
can prevent shear on the discontinuities. Therefore to allow the full development of the
failure, kinematic release within model 12 was ensured by the reduction of the
penalties assigned to the discontinuities, during a post-failure stage (after the rise in
pore pressure had instigated failure).

Results from groundwater modelling


In order to get the rock-bridge to fail as a result of an increase in pore pressure; the
phreatic surface had to be assigned so that, in an unrealistic manner, elements down
the slope face were saturated. Failure did not occur when the phreatic surface is set
back from the slope face (as presented in Figure 6.13a) due to insufficient saturation
within the basal-plane, (this was also the case in Delabole model 11). As in Chapter 4,
it is recommended that a finer mesh density should be used to allow the simulation of a
more realistic curved phreatic surface (although this would increase simulation
runtime).

Figure 6.13: Saturation of base-plane is critical to stability in ELFEN model 12, where limited
pore pressure (N/m2) upon base-plane allows stability (a). Failure occurs in the same model,
when whole length of base-plane is saturated (b).

150
CHAPTER 6: Case study slope 4 (Delabole)

Within the same model the progressive failure across the rock-bridge can be observed.
Figure 6.14 shows the failure of the rock-bridge due to a rise in pore-pressure, causing
subsequent progressive failure and kinematic release. (The video file “Delabole
model 12” on the appended CD-rom is an animation of the failure presented in Figure
6.14).

Figure 6.14: Progressive failure of rock-bridge within ELFEN step-path Delabole model 12, as a
result of a rise in pore pressure.

Importantly this indicated that the degree of pore pressure acting upon the base-plane
is critical to the stability during the Delabole failure. Complete failure was achieved
when the shortah is between 30 and 50% saturated, within a model where the phreatic
surface extends to the slope face, causing pore pressure to act along the whole length
of the base-plane.

Table 6.4 summarises the sensitivity analysis that was conducted using model 12.
Where saturated, the phreatic surface extended to the slope face as depicted in 6.13b.
Instead of conducting a more detailed study, efforts were made to achieve failure within
a model with extended discontinuities, similar to that presented within the latter part of
Section 6.2.

As illustrated in Figure 6.15, model 13, which had extended claylodes, behaved in a
similar way to the step path model, with failure across the rock-bridge initiating from
10s when the final phreatic surface rises to saturate 50% of the shortah. However, as
shown in Figure 6.15d, the block interaction within this model is not as realistic as that
within model 10 outlined at the end of Section 6.2. Consequently it is suggested that
the numerical parameters within model 13, require further consideration.

Despite this, models 12 and 13 have demonstrated the capabilities of ELFEN to


simulate the initiation of the failure at Delabole due to a rise in pore pressure. As a

151
CHAPTER 6: Case study slope 4 (Delabole)

further recommendation, future simulations should investigate the influence of perched


aquifers, as illustrated in Figure 6.1a.

Table 6.4: Results from simulations on Delabole step-path model 12, where the phreatic surface was
raised to different heights, and/or discontinuity frictional strength was altered to cause failure.

Limiting conditions on discontinuities via which


State of Consequent
step-path failure creates a discrete failure block
displacement at FOS assumed

% water filling height of 12 s* in ELFEN**
c (kPa)
shortah
36° 1 0 (dry) Stable >1
36° 1 30 Stable >1
36° 1 50 Ongoing <1
37° 1 50 0.8 m and ongoing <1
38° 1 50 Stable >1

*State of failure is taken at 12s because groundwater rises between 9 and 10s,
therefore any instability due to the rise in groundwater, should have occurred by 12s.

**The FOS within ELFEN is assumed from the degree of ongoing displacement noted;
as discussed in Sections 3.1.1 and 4.5, vertical displacement inevitably occurs in all
simulations when using a dynamic-based code. Where this displacement is minimal

FOS 1.
and a stable state is reached, it has been assumed that ELFEN has predicted a

Figure 6.15: Progressive failure of rock-bridge, as a result of a rise in pore pressure, within ELFEN step-
path Delabole model 13. Full saturation of base-plane and shortah (to 50% height), occurs at 10s
runtime, resulting in fracture of rock-bridge, although further review of block interaction is required.

152
CHAPTER 6: Case study slope 4 (Delabole)

6.4 Summary and conclusions


Throughout Section 6.2, ten different models were introduced in which progressive
failure is simulated via the fracture across rock-bridges, resulting in an active-passive
failure mechanism. Early simulations were especially sensitive, with fracturing before
100% gravity and extensive failure on the initiation of excavation. Subsequently model
boundaries were extended, which improved kinetic stability.

Models 1 and 2 both had excavation objects to restrain the slope face during loading.
Following these simulations it was discovered that using the then current version of
ELFEN, embedded discontinuities must extend through the slope face to achieve
kinematic release (as reported in Section 3.1.1 and Appendix F). It may be a
consequence of this, that that there was no kinematic release in Delabole models 1
and 2.

To promote instability claylodes were extended, and the excavation object removed, to
create model 3 which failed due to gravity loading (with no excavation object); however
again kinematic release does not occur. In this case, this is likely as a result of the
rock-bridge location. As illustrated in Figure 6.4, intense fracturing causes a ‘locking-
up’ behaviour within the failure blocks. Such a problem may be remedied if the
fracturing code could simulate mode II failure, as discussed in Section 2.2.4.

To avoid the intense fracturing, the rock-bridge was altered to encourage tensile failure,
creating a step-path failure. Consequently kinematic release was achieved in models 4
to 6, showing the progression of the model to be dependent on the mesh density.

The subsequent chisel effect, rotation and translation during block interaction, was
achieved through:

1. Implementation of ‘option 166’ (see Appendix A)


2. Prevention of penetration
3. Alteration of rock-bridge location within the shortah.

Experimentation with rock-bridges during this investigation has demonstrated that the
specific location is crucial to ensuring kinematic release and development of a realistic
failure mechanism.

153
CHAPTER 6: Case study slope 4 (Delabole)

Section 6.3 detailed the influence of groundwater within the ELFEN model, primarily
using a simple planar failure simulation to calibrate the numerical parameters and
staged approach. When water was included, this behaved well, although in
comparison to a limit-state approach (RocPlane), the solution in ELFEN was
conservative. This could be due to a disparity in the discontinuity strength applied
within the two approaches; RocPlane considers the shortah as a tension crack and it
appears that it may consider this to have no strength. Consequently future modelling
should investigate this, with re-calibration of an ELFEN model where there is no
strength applied to the shortah.

If there was no discontinuity strength applied to the shortah, one would probably find
that in order to get agreement with RocPlane, the penalties assigned to the ELFEN
base-plane would have to be increased. This may explain why, during the
implementation of groundwater within a step-path model, the penalties had to be
increased to prevent dynamic-related failure of the rock-bridge when the phreatic
surface started to rise.

In addition to the increase in penalties, further modifications had to be made to the


step-path model to be stable in the dry and not in the wet. This is akin to the issues
reported in Section 5.3, where difficulties were experienced during the simulation of a
similar stable-dry, and unstable-wet step-path situation. However in this case, more
appropriate steps were taken to consider and decrease the dynamics from respective
modelling stages. Each of the modifications made have advantages and
disadvantages, and further research is recommended all of which is summarised in
Table 6.5.

Once a successful step-path model was achieved, results were obtained from
groundwater modelling that demonstrated the sensitivity of the Delabole step-path
model to the degree of saturation upon the base-plane. This is very much linked to the
findings reported in Section 4.5.2, with the same conclusions that an accurate, more
curved phreatic surface is required within ELFEN. However, this would demand a fine
mesh density, which would increase runtime. In the case of the current research, the
relatively coarse mesh and consequently simplified phreatic surface provided results
that are accurate enough for the level of analysis required, demonstrating the
successive failure of a rock-bridge due to a rise in pore pressure leading to kinematic
failure within the slope.

154
CHAPTER 6: Case study slope 4 (Delabole)

Table 6.5: The occurrence of premature rock-bridge failure within ELFEN Delabole step-path model, with
prevention methods and recommendations.

Failure
Further
of rock- Prevention by: Disadvantages Modification
recommendations
bridge:
1. Can inhibit the More detailed
10% Rayleigh
failure of the rock- Removed Rayleigh review of loading
damping
bridge damping following process, decrease
the loading stage kinetic energy with
During (replacing
2. Significantly slows (replaced with point the use of
loading alternative
the displacement of damping) conventional point
stage point damping)
sliding blocks damping
Ramp-load as
Review required
opposed to Increases runtime N/A
duration of loading
drop-load
Review of initial
Use of a post-
penalties within a
During failure stage during
Can prevent model where no
initial which penalties are
Increase in kinematic release by strength applied to
rise in reduced to allow
penalties inhibiting shear on the shortah, and find
ground- unhindered
discontinuities limit at which shear
-water comminution and
is prevented due to
deposition
penalties.

Finally groundwater was introduced into a model with a step-path and an extended
fracture network. This more complex model was only studied briefly and with limited
success, consequently it is recommended that there is further research into achieving a
more realistic block interaction.

The following Chapter outlines a back-analysis of a slope failure (case study slope 5 –
Fimiston Open Pit, Kalgoorlie), which is similar to Delabole. Failure occurred on the
same scale and, although not as clearly identifiable, a similar active-passive failure
mechanism could be inferred. However, in case study slope 5 the failure was clearly
stated as initiating in the rock mass (toe region), consequently the main aim in the
following Chapter is to use numerical modelling in a detailed review of mass strength
effects. Rather than using the comprehensive fracture-based code to study the
fracture network and resulting failure mechanism (as occurred throughout this
Chapter), two simpler methods are used in the analysis of rock-bridge strength in a
more direct fashion.

155
CHAPTER 7: Case study slope 5 (Kalgoorlie)

CHAPTER 7: Case Study 5: Modelling of a toe-breakout


mechanism at Kalgoorlie Superpit

7.1 Introduction
This Chapter outlines a back-analysis of rock mass strength at the Fimiston open pit, in
Kalgoorlie, Western Australia. A number of different limit equilibrium solutions, and a
finite method, are used to analyse potential mass strength. Comparisons are made
between each approach, identifying advantages and disadvantages to the modelling
methods.

A slope failure occurred during September 2006 at Kalgoorlie Consolidated Gold Mines
(KCGM) Fimiston open pit (also known as the Superpit). The Superpit is currently
2.5km long, 1km wide and 500m deep; it is orientated with its long axis running north
south, following the gold bearing lode. From the surface at KCGM an approximate
30m deep weathered margin lines the top of the pit; below this the rock mass at KCGM
is competent as a result of the greenstone geology. Principally three lithologies run
through the pit:

 The Golden Mile Dolerite (hosting the gold-bearing lodes);


 Peringa Basalt, and
 the Black Flag Beds.

The failure took place within the Peringa Basalt over several benches at the north
eastern end (Brownhill section) of the pit, as shown in Plate 7.1.

Importantly this failure jeopardised the life of the mine ramp and an alternative ramp
had to be developed. The ramp had been previously closed due to a preceding smaller
failure in July 2005 occurring in the vicinity of Brownhill at -160m level, within the
slightly weathered Peringa Basalt. A back-analysis of the more recent ‘Brownhill
failure’ is of interest to KCGM because a wall steepening project within the pit is
currently under review. As reported in Beer and Morrongiello (2007) this will allow
extraction from deeper levels within the Fimiston open pit.

The following sections detail the key aims, creation of a model, methods of analyses
used and overview of results which were presented to KCGM.

156
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Plate 7.1: (a) Panoramic photograph (looking south) of the Fimiston open pit, Kalgoorlie,
highlighting the failure area; (b) enlarged view of the failed zone. (c) View from the west (looking
north east), with delineation of failed zone.

7.1.1 Key aims of this investigation


The purpose of this investigation is to use back-analysis of the failure as a tool to
understand the potential rock mass strength of the Peringa Basalt. However, when a
cross section of the failed area is studied, as presented in Figure 7.1, one can see that
the failure mechanism involved fracturing (toe-breakout) through a rock-bridge of only
approximately 6m depth. Consequently a back-analysis is greatly affected by strength
of the rock-bridge which represents a very limited portion of the Peringa Basalt
proximal to the face, and therefore liable to have endured blast damage. The back-
analysed strength of the rock-bridge will therefore be somewhat speculative.

Data gained from these studies, can be compared against those presented by Pells
Sullivan Meynink Pty Ltd. in 2004, who used RocLab (Rocscience, 2008) to give rock
mass properties of the Peringa Basalt, with consideration of blasting damage via a
disturbance factor (D) of 1.

157
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.1.2 Creating a model


When an east-west cross section of the failed section is studied (49350m north), as
presented in Figure 7.1, it can be seen that with the structural model that is available,
one can assume that the stability of Brownhill is predominantly a function of the
strengths of the intact mass in proximity to the slope face which forms a rock-bridge,
and the strength of two discontinuities:

1. The Oroya Hanging Wall Fault (OHWF),


2. And the Subvertical Fault (SF).

Figure 7.2: East-west cross section through Brownhill (taken at 49350m north); illustrating the failed
areas, lithologies and major discontinuities.

The geometry formed by these discontinuities, could be compared to an active passive-


type failure (introduced in Section 2.2.5), leaving a ‘chair’-shaped failure scar; although,
in this case analysis was not performed to identify the two separate wedges.

The major underlying fault contributed to the failure mechanism, possibly placing an
increased degree of strain on the overall slope as suggested by
Hewson (2007, personal communication); although there are no advised limits to the
strength of this discontinuity. Instead KCGM provided limits for the frictional strength
and cohesive strength for only two of the discontinuities (the OHWF and the SF), as
listed within the next section in Table 7.3. In addition, a prominent foliation dipping into
the pit was noted during the site visit, in the bench behind the failure (within the Peringa
Basalt). Also Beer and Morrongiello (2007) state that it was reported by a consultant in

158
CHAPTER 7: Case study slope 5 (Kalgoorlie)

2000 that there is a long history of instability limited to only the north end of the east
wall where there is 65 foliation. It is assumed that this is the same foliation that was
noted, on the bench at the top of the failed section, during the site visit.

It is important to note at this stage that all analyses have been performed with the
understanding that there is no influence of groundwater within the failure area. This
has been assumed because the Fimiston open pit has limited groundwater problems
due to uniquely good drainage as a result of the numerous old mine workings within the
floors and walls of the pit, which previously worked the ore body.

7.1.3 Methods of Analysis


Five different limit state models were developed and two different finite element models
were used to interrogate the strength of the rock-bridge given the conditions on the
discontinuities suggested by KCGM, which are listed in several of the Tables (7.3, 7.8,
7.10 and 7.11). Table 7.1 lists the advantages of each of the limit equilibrium models,
and Table 7.2 discusses the finite element models that were developed. Importantly
the basic construction and consequent disadvantages of each of model is stated; this
exposes aspects which cannot be considered within some of the modelling methods.

159
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Table 7.1: Basic construction of the five different limit equilibrium models used for the analysis of cohesive strength of the rock-bridge; with advantages and
disadvantages listed for each technique.

Model
Method of rock-bridge cohesive
used for Construction Disadvantages related to construction Advantages to technique
strength analysis
analysis
Weight of sliding mass Cannot consider:
 Any component of strength on SF;
Polygon of
summed from slices in Jacob  Quickest and simplest
 c on OHWF;
Forces
Uses an equivalent bolt tension for
  or σt on rock-bridge.
(Limit Model 1, with trigonometry to method of analysis.
rock-bridge load capacity.
Equilibrium calculate consequent
Method) resisting force required.
 Involves the simplification of the geometry
Jacob
 Once simplified geometry is
into a maximum of 21 equal slices.
 A large horizontal driving force is imparted
Model 1 Back-analysis of strength on a
Uses simplified profile divided
(Limit discontinuity that represents the setup, sensitivity analysis is
into slices on the rest of the sliding mass, by final 3
Equilibrium failure through the rock-bridge. quick.
slices.
 Cannot consider σt
Method)

 Involves the simplification of the geometry


Jacob
As in Jacob Model 1;
Model 2
however final 3 slices are into a maximum of 21 equal slices.
 Cannot consider σt
(Limit As in Jacob Model 1 As in Jacob Model 1
removed eliminating high
 Cannot consider any strength on SF.
Equilibrium
horizontal driving force.
Method)
Plane Weight of sliding mass is
Equivalent cohesive strength of rock- Cannot consider:
taken and single plane failure
 Any component of strength on SF;
Failure
bridge is taken from degree of tension
 Quick Method
  on OHWF;
Model 1 is constructed with basal
required upon a bolt inclined at the
  or σt on rock-bridge.
(Limit plane representing OHWF
same angle as the discontinuity that
Equilibrium and tension crack
represented the rock-bridge.
Method) representing SF
Plane
 As in Plane failure model 1;
Failure Equivalent cohesive strength of rock-
 Also cannot consider cohesive strength on
Model 2 bridge is taken from degree of
As in Plane Failure Model 1 As in Plane Failure Model 1
(Limit cohesive force required upon a base
OHWF
Equilibrium plane (OHWF) to resist failure
Method)
160

160
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Table 7.2: Basic construction of the two different finite element models used for the
analysis of cohesive strength of the rock-bridge; with advantages and disadvantages
listed for each technique.

Method of rock-
Model
bridge cohesive Disadvantages related Advantages to
used for Construction
strength to construction technique
analysis
analysis

 Can consider
 Increased numerical
Accurate profile
Back-analysis of
can be tensile strength
strength on a parameters;
 Potential influence of
constructed of rock-bridge,
Phase2 discontinuity that
with embedded unloading via
Model 1 represents the unloading technique
discontinuities excavation and
failure through and proximity of
between the in-situ stress
the rock-bridge. model boundaries.
elements regime.

As in Phase2 model
2
As in Phase model 1. 1. Also:
Analysis of rock- also:  Potentially most
bridge strength  Influence of mesh reliable
directly allowing density around rock- technique,
the failure of bridge is significant; allowing failure
Phase2 As in Phase2 rock-bridge via  Increased influence to take
Model 2 model 1. straining of of unloading undefined path;
elements and technique,  SSR technique
consequent consequent more is possible
yielding on appropriate staged allowing
OHWF and SF excavation amounts automated
to increased runtime. detection of limit
point.

7.1.4 Overview of results


Failure of the rock-bridge was found to be impossible if the strengths of the
discontinuities are set to their maximum, within the range given by KCGM. Instead
back-analyses were conducted using the minimum discontinuity strengths and in some
cases zero strength on the SF. Consequent findings were that:

 The limit equilibrium models demonstrated that the potential cohesive strength
of the rock-bridge could be between 200-400kPa, with the exception of one
model (Jacob Model 1), which gave very high and potentially anomalous data
unless there was a degree of strength applied to the SF.
 Phase2 models showed that the cohesive strength of the rock-bridge could
perhaps be slightly higher, resulting in strengths of up to approximately
500kPa.

The use of Phase2 gave various benefits and set-backs, with modifications to models
necessary, in order to give reliable data. Importantly it was found that the in-situ stress
regime had a minimal influence on rock-bridge strength.
161
CHAPTER 7: Case study slope 5 (Kalgoorlie)

The influence of the major underlying fault remains untested. Shear strength data for
this discontinuity would be required; also a greater degree of confidence is necessary
for the strengths of the OHWF and the SF. More importantly perhaps, increased
structural detail is required within the toe-region to aid development of a more
comprehensive numerical model. This could potentially determine the influence of the
adversely orientated foliation, and increase understanding of the rock-bridge strength.

7.2 Limit State Analyses of rock-bridge strength


When a cross section of the failed section is considered, as illustrated in Figure 7.1, the
weight of the rock volume that the rock-bridge is resisting is relatively small
(approximately 16 tonnes per metre within the plane of the section). Consequently
when this is considered in back-analysis, it could be anticipated that the cohesive
strength of the rock-bridge is likely to be most important in providing resistance to the
failure, as opposed to mass frictional strength.

A method of evaluating the magnitude of the cohesive strength of the rock-bridge is to


analyse this problem as a limit-equilibrium, with the assumption that the toe-breakout
occurred on a discontinuity of which the conditions will be determined. The approach
taken was to construct the problem in Jacob.xls (Pine, 2006a). This involved
simplification of the geometry so that the geometry could be divided into no more than
twenty one slices, as shown in Figure 7.2. From this an accurate total weight of the
sliding block in question can be taken and used in the first part of this back analysis,
the creation of a polygon of forces.

Figure 7.2: Annotated cross section through Brownhill, illustrating the actual and simplified
profile, which can be split into 21 slices for analysis in Jacob.

162
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.2.1 Polygon of Forces


In the initial analysis the failed section of Brownhill can be interrogated as a single
mass sliding on a single discontinuity (represented by the Oroya Hanging Wall Fault).
The cohesive strength of the rock-bridge can be back-calculated via determining the
equivalent tensile force, which is inputted via a notional rock-bolt, to prevent this mass
from sliding. Note that the bolt is inclined at the same angle as the plane that
represents the toe break-out. This technique is depicted in Figure 7.3.

Figure 7.3: Polygon of forces approach to analysing the cohesive


strength of the rock-bridge required to prevent the mass from sliding.

failure; T = cohesive strength of rock bolt representing the rock-bridge; = inclination of OHWF;
(Where W = weight of sliding mass; N = normal force on OHWF; f = frictional force on OHWF;

 = strength of OHWF and = Inclination of rock bolt representing the rock-bridge).

This approach shows that the minimum cohesive strength for the rock-bridge is
approximately 187kPa when the following assumptions are taken:

 Total Weight of Slices (W) = 16269 kN/m


 Subvertical Fault has no strength
 OHWF  = 24°; c = 0kPa; Dip () = 28°
 Length of discontinuity representing line of toe-breakout = 6m; dip ()= 10°

When the frictional strength of the OHWF is assumed to be at its lowest limit of 20
(within the boundaries advised by KCGM), then the required cohesive strength of the
rock bridge increases to 379 kPa.

163
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.2.2 Multiple slices approach – Jacob


Jacob Model 1 – default (high horizontal forces in last three slices)
Using the Excel-based spreadsheet Jacob created by Pine (2006a) it is possible to
perform sensitivity analysis on the strength of the rock-bridge using the profile
presented in Figure 7.2. Importantly, within the following results (from Jacob Model 1),
it was noted by Pine (2008, personal communication), that the slices representing the
material above the Subvertical Fault, impart an unusually large horizontal driving force
promoting instability of the slope as illustrated in Figure 7.4. This could be realistic as
the dip of the SF mapped within the Vulcan section that was taken, is most
unfavourable concerning additional horizontal force application. It is important
however, to investigate the influence of this, which is why two Jacob models have been
created, as shown in Figure 7.4.

Figure 7.4: Profile of Brownhill Jacob Models 1 and 2, with annotation illustrating the
slices from which the high horizontal force originates in Jacob Model 1.

Table 7.3 illustrates the results from a sensitivity study analysing the rock bridge
strength and the influence of the variability of the given variables within Jacob Model 1
(which has the high horizontal force).

This study is based around the following cases:

 Case 1 determines the FOS when the median values for discontinuity strength,
(within range of strengths on discontinuities given by KCGM), are applied and
no additional strength in the rock-bridge.
 Case 2 determines the influence of the cohesion applied to the OHWF and the
subsequent cohesion of rock-bridge required to achieve a FOS of close to 1;

164
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Cases 3 to 8 give the relevant cohesion required on rock-bridge, to achieve a FOS of


close to 1 when:

3. The same conditions are applied as those within the polygon of forces model;
4. The lowest values for discontinuity strength, (within the range given by KCGM),
for OHWF and SF;
5. Friction also for the rock-bridge and minimum shear strength applied to the
OHWF and SF;
6. Zero strength is applied to the SF and the OHWF has 20°  only;
7. There is  as well as c applied to the rock-bridge and no shear strength applied
to SF;
8. There is  as well as c applied to the rock-bridge the OHWF has 24°  and no
shear strength on SF.

Table 7.3: Results of a sensitivity study on the strength of the rock mass that forms the rock-
bridge in Jacob Model 1; values emboldened highlight what values have been altered from the
preceding case.

Case 
Oroya Hanging
Rock Bridge Subvertical Fault
Wall Fault
Conditions Conditions
Conditions

Ranges   = 20-24˚  = 24-30˚


KCGM c = 0-150 kPa c = 150-300 kPa Resulting FOS
?
(Janbu)

 = 0˚  = 22˚  = 27˚
c = 0 kPa c = 75 kPa c = 225 kPa
1 1.60

 = 0˚  = 22˚  = 27˚
c = 0 kPa c = 0 kPa c = 225 kPa
2 1.19

 = 0˚  = 24˚  = 0˚
c = 950 kPa c = 0 kPa c = 0 kPa
3 1.02

 = 0˚  = 20˚  = 24˚
c = 110 kPa c = 0 kPa c = 150 kPa
4 1.00

 = 40˚  = 20˚  = 24˚


c = 0 kPa c = 0 kPa c = 150 kPa
5 1.03

 = 0˚  = 20˚  = 0˚
c = 1100 kPa c = 0 kPa c = 0 kPa
6 1.03

 = 40˚  = 20˚  = 0˚
c = 1150 kPa c = 75 kPa c = 0 kPa
7 1.02

 = 40˚  = 24˚  = 0˚
c = 1000 kPa c = 0 kPa c = 0 kPa
8 1.02

165
CHAPTER 7: Case study slope 5 (Kalgoorlie)

From this sensitivity study it was found that:

 The cohesive strength on the OHWF has a substantial influence on the FOS
and therefore the potential cohesive strength for the rock-bridge.
 Jacob Model 1 requires a significantly higher (750kPa) cohesion for the rock-
bridge than that predicted by the same conditions within the Polygon of Forces.
 The condition of the rock-bridge at the toe is problematic; it is unknown whether
the rock mass is significantly damaged so that it is comprised of small, almost
discrete blocks, (in which case it would be mainly frictional strength).
Alternatively it may be more intact, in which case the cohesive strength would
be the main component within the rock mass. This appears to be critical when
the minimum strength conditions are applied to the sub-vertical fault, as in
cases 4 and 5. Only either a small degree of cohesive strength, or a
reasonable amount of frictional strength can be applied on the rock-bridge in
these cases

Jacob model 2 - Removal of driving force from final three slices


To counteract the high horizontal forces, the three final slices that represent the
material above the SF were removed. Their weight was added to the remaining 17
slices, by increasing the unit weights.

The resulting cohesive strength of the rock-bridge, required to get a FOS of 1 when the
lowest strengths were taken, was 300kPa. A similar sensitivity study as previously
conducted was analysed in the model with the deactivated final slices. The same case
numbers are used as in the previous models.

The same conclusions as previously can be made; however importantly case 8


illustrates that a much lower cohesion (800kPa lower), is required on the rock-bridge to
prevent failure when the lowest strength is assumed (within the range given by KCGM),
in comparison to the previous modelling approach. This is most likely due to the
elimination of the high horizontal force, by removing the slices in Jacob model 2.

Case 6 can be compared directly back to the result from the polygon of forces analysis,
interestingly demonstrating exactly the same result with 187 kPa of cohesion required
on the plane representing the rock-bridge, to prevent failure. The similarity within these
two methods is most probably due to the lack of consideration of strength for the SF
within both approaches.

166
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Table 7.4: Results of a sensitivity study on the strength of the rock mass that forms the rock-bridge
in Jacob Model 2; values emboldened aid the viewer to see what values have been altered from the
preceding case.

Rock Bridge Oroya Hanging Wall Fault Resulting FOS


Case
Conditions Conditions (Janbu)

 = 0˚  = 22˚
c = 0 kPa c = 75 kPa
1a* 1.01

 = 0˚  = 22˚
c = 250 kPa c = 0 kPa
2a* 1.01

 = 0˚  = 20˚
c = 300 kPa c = 0 kPa
6 1.00

 = 40˚  = 20˚
c = 170 kPa c = 0 kPa
7 1.00

 = 0˚  = 24˚
c = 187 kPa c = 0 kPa
3 1.00

 = 40˚  = 24˚
c = 25 kPa c = 0 kPa
8 1.00

*Note that these are different to cases 1 and 2 in Jacob Model 1, as with the Jacob
Model 2 analysis no strength can be assigned to the SF.

Summary
The multiple slice limit equilibrium approach has enabled a more detailed sensitivity
study of the influence of input parameters. The range of cohesion for the rock-bridge
analysed by the multiple slice limit method, is from 110 to 650kPa.

Jacob Model 2 can be used to show that the likely cohesion of the rock bridge is
potentially lower, around 200kPa. Also the model closely agrees with the polygon of
forces method, however as previously stated this model (and the Polygon of Forces) do
not take the strength of SF into account.

The most significant questions that should be taken from this sensitivity study, of the
influences of each of the parameters within these analyses, are as follows:

1. How reliable are the strength conditions on the OHWF and SF?
2. How confident can one be of the high horizontal force (which is a function of dip
of the SF), being imposed on the sliding mass as a result of the material above
the SF?

167
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.2.3 Single block – plane failure model


Another method to back-analyse the cohesive strength of the rock-bridge using a limit
equilibrium approach, is to consider the sliding mass as a planar failure, with a single
block sliding having similar weight and dimensions on a single plane (representing the
OHWF), releasing via a tension crack (representing the SF). The software used was
the Excel-based spreadsheet Plane_failure.xls, developed by Pine (2006b).
Fundamentally the cohesive strength of the rock bridge can then be calculated via
analysing:

1. The tension required in an equivalent bolt to prevent the mass from failing, (bolt
is inclined at the same angle as the discontinuity representing the toe-
breakout); or
2. The cohesive strength required on the base plane to prevent the mass from
failing (assuming that the base plane/ OHWF is purely frictional).

The profile used and the two techniques are presented in Figure 7.5.

Figure 7.5: Plane_failure.xls (Pine, 2006b) model for Brownhill,


where the deficit in resistance on the base plane is equated to
the cohesive strength of the rock-bridge via either the tension
required on a bolt, or the cohesive strength of the base plane.

Table 7.5 lists the results from a sensitivity study on the cohesive strength of the rock
bridge when calculated using the first approach listed above. Again the same cases
are preserved as those used in the previous analyses. Consequently case 6 considers
the lowest strength within the range, and case 3 uses the higher frictional strength
within the range. Cases 10 and 11 are used to illustrate the large influence of cohesive
strength applied to the base plane (OHWF).
168
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Table 7.5: Results from sensitivity study using Plane_Failure.xls (Pine, 2006b), with resulting cohesive
properties of the rock bridge, when calculated from the tension required in a rock bolt.

Oroya Hanging Subsequent cohesive


Tension required on Rock-bolt
Case Wall Fault strength of 6m Rock
for FOS = 1 (kPa)
Conditions Bridge (kPa)

 = 22
c = 75 kPa
1 - -

 = 20
c = 0 kPa
6 2300 383

 = 24
c = 0 kPa
3 1150 192

 = 22
c = 10 kPa
10 1300 217

 = 20
c = 50 kPa
11 300 50

Table 7.6 contains the results from the second approach listed above. Note that
cohesive strength of the OHWF cannot be considered, as in this model this aspect is
instead used to calculate the cohesion of the equivalent rock-bridge. Consequently
case 9 incorporates the minimum frictional strength of the OHWF, and case 3 the
maximum.

Table 7.6: Results from sensitivity study using Plane_Failure.xls (Pine, 2006b), with resulting
cohesive properties of rock bridge, calculated from the cohesive strength required on the base plane
to prevent the mass from failing.

Subsequent
Oroya Hanging Cohesion required on Resulting
cohesive strength of
Case Wall Fault base-plane for cohesive force
6m Rock Bridge
Conditions FOS = 1 (kPa) (kPa)
(kPa)

 = 20
c = 0 kPa
9 58 2457 410

 = 24
c = 0 kPa
3 30 1271 212

169
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.2.4 Discussion
The results from the limit equilibrium methods give a wide range of potential cohesive
strengths for the rock-bridge, with an inevitably strong dependency on strengths
applied to OHWF and SF (where possible).

 The polygon of forces method, demonstrates likely cohesive strength of rock-


bridge is between approximately 200-400kPa dependent on the  of the OHWF;
 Jacob Model 1 consistently predicts a much higher cohesive strength for the
rock-bridge (≥ 950kPa), unless there is a degree of strength on the SF, in which
case the cohesion of rock-bridge = 110kPa, (alone this model accounts for the
horizontal force created by the inclined SF);
 Jacob Model 2 can be used to show that the likely cohesive strength for the
rock-bridge is potentially lower, around 200-300kPa dependent on the  of the
OHWF;
 Both planar failure models agree to a reasonable extent, with a cohesive
strength ranging from 200-400kPa, dependent on the  of the OHWF.

Table 7.7: Summary of results from the sensitivity study of rock-bridge cohesive strength, using
different limit equilibrium models.

 of rock mass
Cohesive Oroya
Subvertical Limit
strength of Hanging
Rock-Bridge ()
Case comprising Fault equilibrium
rock-bridge Wall Fault
Conditions model
(kPa) Conditions

 = 0
c = 0 kPa
1100 Jacob Model 1

Plane Failure
410 -
Model 2

 = 20
0 c = 0 kPa
6 Plane Failure
383 -
Model 1
Polygon of
379 -
Forces
300 - Jacob Model 2

 = 0
c = 0 kPa
950 Jacob Model 1

Plane Failure
212 -
Model 1
 = 24
c = 0 kPa
3 0 Plane Failure
192 -
Model 2
Polygon of
187 -
Forces
187 - Jacob Model 2

 = 24  = 0
c = 0 kPa c = 0 kPa
8 1000 40 Jacob Model 1

 = 20  = 24
c = 0 kPa c = 150 kPa
4 110 40 Jacob Model 1

170
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.3 Phase2 Modelling


Via the application of Phase2 (Rocscience, 2008), one can consider a number of issues
(such as the in-situ stress, sequence of excavation and additional strength properties)
that may influence the strength of the rock-bridge. A detailed list of benefits of the
hybrid approach within Phase2 is given in Table 2.2.

Two methods can be used to represent the rock-bridge in Phase2:

Model 1: The model can be constructed with a discontinuity to represent the rock
bridge (as was the case within the limit equilibrium models)
Model 2: Alternatively a model without this discontinuity can be created allowing
the analysis of strain within the finite elements constituting the rock-
bridge through which toe-breakout occurred.

Both these methods have been used and the results are presented below. Importantly
prior to a detailed study it is valuable to consider aspects that may influence the results:

 The proximity of boundaries


 The effect of unloading via a staged excavation process
 The tensile strength of both the mass and discontinuities
 The stiffness properties of discontinuities
 The consequence of a non-uniform in-situ stress regime.

An additional aspect that could importantly influence only the second Phase2 model is
mesh density, which is likely to have an influence on the apparent strain that occurs
through the rock bridge.

Firstly a relatively simple small single-staged model was created as a progression from
the limit equilibrium analysis. A detailed study was conducted using the model
illustrated in Figure 7.6a; however it was later found that this had inherent inaccuracies
due to the close proximity of the boundaries (significantly lower cohesive strengths
were required on the rock bridge to resist failure). Consequently the model illustrated
in Figure 7.6b was constructed.

Also the process by which the slope is formed could have quite a significant influence
on the rock-bridge strength that is back-analysed. Potential stages of excavation are
illustrated in Figure 7.6a, showing a four stage excavation. Alternatively a single-
staged simulation can occur where the slope exists from the beginning of the
simulation (no staged excavation).

171
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Figure 7.6: (a) Primary Phase2 Model 1, with degree of straining presented; (b) same model but
extended boundaries causing a reduction in straining in the region close to the Subvertical Fault.

7.3.1 Phase2 Model 1: Discontinuity Model


It was found that in the case of a frictionless rock-bridge, the influence of staged
excavation was minimal. Therefore to minimise runtimes, the decision was taken to
proceed with the analysis using the un-staged model (with extended boundaries). In
addition occasional checks were made, running the same simulation within the staged
model, to confirm the minimal influence of a staged excavation procedure. In particular
when a five staged excavation sequence was used rather than a single staged model;
it was found that the limit of stability for cohesion of the rock-bridge was:

172
CHAPTER 7: Case study slope 5 (Kalgoorlie)

 10kPa lower for case 4 (minimum strength on OHWF and SF).


 10kPa higher for case 6 (minimum strength on OHWF and no strength on SF).

Consequently it could be assumed that all values within Table 7.8 are subject to a
 10kPa variance; this is insignificant in the scale of rock-strength therefore in this case
the staged unloading sequence has little influence.

The results from an analysis of Phase2 Model 1 are presented in Table 7.8; note that in
addition to the cases from the previous studies, cases 12-14 have been introduced.
Cases 4 and 5 show that if the minimum strength (within the range given by KCGM)
was assumed for the SF then the strength of the rock-bridge should come from either
independently frictional or cohesive strength, or alternatively a very small amount of
both. Case 8 develops this but with zero strength on the SF.

Cases 4 and 13 demonstrate that in this particular model tensile strength has no
influence to the stability of Brownhill. Case 6 is a direct comparison to the
Plane_failure.xls model reported in Section 7.2.3. This illustrates that both the models
created in Plane_failure.xls and Jacob Model 2, perhaps slightly under-estimated the
cohesive strength for the rock-bridge, and in fact a higher (100kPa greater) cohesive
strength is predicted for the rock-bridge when there is no strength assumed for the SF
and only frictional strength on the OHWF.

Importantly from this set of Phase2 analyses, the following conclusions can be drawn:

1. The lowest amount of cohesive strength that can be assumed for the rock-
bridge is near to zero if a reasonable amount of frictional strength (40°) is
assigned to the rock-bridge.
2. The highest amount of cohesive strength that can be assumed for the rock-
bridge is 480kPa, when there is no frictional strength assigned to the rock-
bridge and the SF has no strength properties.
3. The most likely amount of cohesion for the rock-bridge is 150kPa, when there is
zero frictional strength assigned to the rock-bridge and the minimum strengths
from the range (given by KCGM) for the respective discontinuities are used.

The following section presents the results from the second Phase2 Model, providing a
final method of analysis for rock-bridge strength in the toe of Brownhill. In addition, an
aspect that was not considered during the analysis of Phase2 Model 1 is in-situ stress.
In the case of Phase2 Model 1 and the initial Phase2 Model 2 analysis, the k ratio
remained at the default within Phase2 (σh = σv). The influence imposing a higher and
more realistic k ratio is analysed in the later part of the following section.

173
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Table 7.8: Phase2 (Model 1) study of rock-bridge strength. Note that the analysis was performed
sequentially, with the parameter which has been altered between each case highlighted in bold.

Oroya
Subvertical
Rock Bridge Hanging Wall Percentage of yield along
Fault
Conditions Fault respective discontinuities*
Conditions
Conditions
c = 0 to c = 150 to
KCGM Rock-
 = 20 to 24  = 24 to 30
Case ? 150 kPa 300 kPa SF OHWF
Ranges: bridge

 = 0  = 20  = 24
c = 165 kPa c = 0 kPa c = 150 kPa
Minimum 75 100 100
4 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0˚  = 20  = 24
c = 160 kPa c = 0 kPa c = 150 kPa
Minimum 100 100 100
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0˚  = 24  = 24
c = 45 kPa c = 0 kPa c = 150 kPa
Minimum 75 100 100
12 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0˚  = 24  = 24
c = 40 kPa c = 0 kPa c = 150 kPa
Minimum 100 100 100
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 40  = 20  = 24


c = 50 kPa c = 0 kPa c = 150 kPa
Minimum 0 100 0
σt = 0 kPa σt = 0 kPa σt = 0 kPa
5
 = 40  = 20  = 24
c = 0 kPa c = 0 kPa c = 150 kPa
Minimum 0 100 20
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0˚  = 20  = 24
c = 160 kPa c = 0 kPa c = 150 kPa
Minimum 75 100 100
13 σt = 100 kPa σt = 0 kPa σt = 0 kPa

 = 0˚  = 20  = 24
c = 165 kPa c = 0 kPa c = 150 kPa
Minimum 100 100 100
σt = 100 kPa σt = 0 kPa σt = 0 kPa

 = 40  = 20  = 0
c = 65 kPa c = 0 kPa c = 0 kPa
Below
100 100 75
minimum
7 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 40  = 20  = 0
c = 60 kPa c = 0 kPa c = 0 kPa
Below
100 100 100
minimum
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 0
c = 485 kPa c = 0 kPa c = 0 kPa
Below
100 100 0
minimum
6 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 0
c = 480 kPa c = 0 kPa c = 0 kPa
Below
100 100 100
minimum
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 40  = 24  = 0
c = 0 kPa c = 0 kPa c = 0 kPa
Below
14 100 100 50
minimum
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 24  = 0
c = 370 kPa c = 0 kPa c = 0 kPa
Below
100 100 75
minimum
3 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 24  = 0
c = 365 kPa c = 0 kPa c = 0 kPa
Below
100 100 100
minimum
σt = 0 kPa σt = 0 kPa σt = 0 kPa
* As illustrated later in Figure 7.7, the degree of yielding upon embedded joints was
used to indicate the state of stability predicted by Phase2.

174
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.3.2 Phase2 Model 2: Non-discontinuity model (breakout through


rock mass)
Primarily it was important to assess the influence of mesh density; a model was
constructed using the default (coarse) mesh density, this gave a considerably lower
strength than a model with a finer mesh, as shown in Table 7.9. This was within a
model with no strength applied to the SF and a  of 20° applied to the OHWF (c and σt
= 0); in addition the models from which the results were obtained are illustrated in
Figure 7.7.

Table 7.9: Influence of mesh density within (un-staged) Phase2 Model 2 of Brownhill failure (no strength
applied to the SF and purely frictional strength of 20° applied to OHWF).

Back-analysed cohesive strength Percentage of yield along


of mass (kPa) within: respective discontinuities: Consequent
Coarsely Fine meshed FOS assumed
SF OHWF
meshed model model
50 110 100 75 >1
45 105 100 100 <1

Another important aspect is to clarify the influence of unloading via staging excavation,
sequentially excavating layers 1 to 4 as illustrated Figure 7.7a, within a finely-meshed
non-discontinuity model. This was also significant, with cohesion being almost
100kPa (90%) lower that that listed in Table 7.9, when a staged excavation process
was implemented:
 For a FOS >1 rock-bridge c = 20kPa, whereas 15kPa gave a FOS <1, in the
same situation (zero strength applied to SF and only a 20°  applied to OHWF).

Figure 7.7: Comparison between results from same Phase2 Model 2, but with different mesh
densities; strain is contoured on same interval for both (a) and (b). (Focussed SSR search area is
illustrated in (a) for the purpose of SSR modelling outlined later in this section).

175
CHAPTER 7: Case study slope 5 (Kalgoorlie)

The results from these preliminary simulations, illustrates the importance of setting a
fine mesh density within the toe-breakout region and also performing a staged analysis
within a non-discontinuity model. However both these modifications considerably
increase the model runtime.

As this model involves straining within the rock mass, rather than failure along a
discrete feature (as was the case in Model 1), the shear strength reduction (SSR)
analysis technique can be used to minimise the time spent finding the limit of stability.
However the SSR method is limited, as the user cannot define the SSR to occur upon
the c or  of the rock-bridge independently; instead both are reduced by the same
shear strength reduction factor (SRF), as was shown within Equation 2.4. Although it is
possible within Phase2 to apply a tensile strength to the mass, which can be excluded
from the SSR; this allows one to independently analyse the influence of tensile strength
on rock-bridge strength, as in Phase2 Model 1.

It was found that the SSR search area must be defined so that it is focussed to
decrease the shear strength within only the toe region, as indicated in Figure 7.7a.
This significantly reduces computational effort, with an unlimited model taking
approximately twenty minutes to run whereas a limited model took between five and
ten minutes, on a 2.4GHz single processor with 1GB of RAM. Also more importantly
different results occur when the SSR technique is applied to the whole model and a
subsequent global reduction in shear strength occurs; for instance the limit of stability
for case 6 (within Table 7.10) was 100kPa lower within an unlimited SSR model. This
could be due to the fact that within a model where the SSR is global, it is not just the
rock-bridge that is being down-graded, instead the simulation involves a degradation of
the whole slope, which is unrealistic in terms of the limited area of potential blast
damaged (weakened) rock that the rock-bridge occupies.

Table 7.10 presents the results of the sensitivity of Phase2 Model 2 using the SSR
technique (within a model with extended boundaries, and a five stage excavation
sequence). As with the Phase2 discontinuity model, cases 4 and 13 show that tensile
strength does not influence the strength of the rock-bridge. When the other results are
compared back to the discontinuity model one can see that the discontinuity model
could have potentially been behaving in a conservative manner, revealing higher
strength for the rock-bridge, (between 60 and 80kPa higher), than in the non-
discontinuity model.

176
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Table 7.10: Results of a sensitivity study on the strength of the rock mass that forms the rock-bridge in
Phase2 Model 2 (using a default k ratio of 1).

Percentage of
Oroya Hanging
Rock Mass Subvertical Fault yield along
Wall Fault
Conditions Conditions respective
Conditions
discontinuities

 = 20-24  = 24-30
KCGM c = 0-150 kPa c = 150-300 kPa
Case ? SF OHWF
Ranges:

 = 0  = 20  = 24
c = 111 kPa c = 0 kPa c = 150 kPa
Minimum 50 75
4 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 24
c = 100 kPa c = 0 kPa c = 150 kPa
Minimum 100 100
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 24
c = 111 kPa c = 0 kPa c = 150 kPa
Minimum 100 100
13 σt = 100 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 24
c = 100 kPa c = 0 kPa c = 150 kPa
Minimum 100 100
σt = 100 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 0
c = 444 kPa c = 0 kPa c = 0 kPa
Below
100 75
Minimum
6 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 0
c = 400 kPa c = 0 kPa c = 0 kPa
Below
100 100
Minimum
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 24  = 0
c = 266 kPa c = 0 kPa c = 0 kPa
Below
100 75
Minimum
3 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 24  = 0
c = 250 kPa c = 0 kPa c = 0 kPa
Below
100 100
Minimum
σt = 0 kPa σt = 0 kPa σt = 0 kPa

Influence of in-situ stress


The data presented in Table 7.10 should be considered as the most realistic out of all
of the limit equilibrium and finite element analyses carried out, considering there is no
pre-defined weakness along which the toe-breakout must occur. However the final
parameter to test, is the influence of the in-situ stress regime. Data can be taken from
a CSIRO HI Cell Overcoring stress study within the Peringa Basalt, which was
conducted by Pascoe in 1999, as reported in a Microsoft© Excel spreadsheet
(MtC+Fimiston RSM.xls) within the stress analysis folder of rock properties on the
internal KCGM geotechnical database. The largest magnitudes of stress reported were
within the “Chaffers Shaft” (at 194 m below the surface) with:

 The major principal stress (σ1) as 18.2MPa in the horizontal north-south


direction;
 The intermediate principal stress (σ2) as 12.3MPa also in the horizontal east
west direction, and
 The minor principal stress (σ3) as 5.3MPa in the vertical.
177
CHAPTER 7: Case study slope 5 (Kalgoorlie)

All models used for the Brownhill analysis during this research are taken from an east-
west section; therefore the major in-situ stress is out-of-plane. However the in-plane
intermediate principal stress is still significant, forming a k ratio of 2.32 (σ2/σ3). This
could be causing a considerable degree of damage in the slope caused by the release
of high in-plane stresses, which would consequently decrease the strength required to
inhibit failure. Consequently a model was created, as illustrated in Figure 7.8, with a k
ratio of 2.32 the results from which are presented in Table 7.11.

Figure 7.8 illustrates the progressive yielding of the Oroya Hanging Wall Fault (OHWF)
occurring once excavation of the final stage had occurred, with 100% yield when the
strength within the rock-bridge had been reduced by a factor of 2.2. The initial strength
of the mass for this model was purely cohesive at 800kPa; a reduction of 2.2 means
that the yield point must be at approximately 364kPa (as presented in case 6 within
Table 7.11).

Figure 7.8: Development of shear strain (in case 6 within Table 7.11) along
rock-bridge and consequent yielding upon discontinuities; due to excavation
and strength reduction within the rock-bridge (SRF = shear strength reduction
factor).

178
CHAPTER 7: Case study slope 5 (Kalgoorlie)

Table 7.11: Analysis of the strength of the rock mass that forms the rock-bridge in a Phase2 Model
2,with a k ratio of 2.32.

Percentage of
Oroya Hanging
Rock Mass Subvertical Fault yield along
Wall Fault
Conditions Conditions respective
Conditions
discontinuities

 = 20-24  = 24-30
KCGM c = 0-150 kPa c = 150-300 kPa
Case ? SF OHWF
Ranges:

 = 0  = 20  = 24
c = 111 kPa c = 0 kPa c = 150 kPa
Minimum 50 75
4 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 24
c = 100 kPa c = 0 kPa c = 150 kPa
Minimum 100 100
σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 0
c = 400 kPa c = 0 kPa c = 0 kPa
Below
100 75
Minimum
6 σt = 0 kPa σt = 0 kPa σt = 0 kPa

 = 0  = 20  = 0
c = 364 kPa c = 0 kPa c = 0 kPa
Below
100 100
Minimum
σt = 0 kPa σt = 0 kPa σt = 0 kPa

Comparisons between the results in Table 7.10 and 7.11 demonstrate that a high in-
situ stress has no influence on the strength of the rock-bridge when the minimum
strength, (within the range suggested by KCGM), is applied to the discontinuities and
the rock-bridge is purely cohesive. However when there is zero strength applied to
Subvertical Fault, the k ratio of 2.32 causes a reduction in the back-analysed cohesive
strength of the rock-bridge (363kPa as opposed to 400kPa). This is a relatively
insignificant difference, indicating the in-situ stress regime has a potentially minimal
influence on the rock-bridge strength. Importantly values of cohesive strength still fall
within the 100 to 450kPa range for the cohesive strength of the rock-bridge, which
seems to be consistent throughout all of the analyses.

179
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.4 Summary and Discussion


Four different modelling and analytical techniques were used, although within these a
further three methods were also employed, to derive mass strength of the rock-bridge.
The first section within this summary compares and discusses the results from each of
these approaches; the second section concludes the findings from this chapter.

7.4.1 Comparisons between modelling approaches


The range of potential discontinuity strengths for the OHWF and SF, suggested by
Hewson (2008, personal communication) is quite large; this caused a significant
variance within the range of back-analysed cohesive strengths for the rock-bridge.
Importantly failure is impossible if the strengths of the discontinuities are set to their
maximum. Therefore as a summary, four situations were considered:

1 and 2 - below the minimum; 3 and 4 - at the minimum discontinuity strengths.

Further detail of each situation is presented in Table 7.12; also the methods which
could be used for each situation are listed. This is because, as discussed in Tables 7.1
and 7.2, some of methods of analyses cannot be used in certain situations. Figure 7.9
illustrates the subsequent results.

Table 7.12: Situations used to compare the different methods of analysis that have been used during the
investigation of rock-bridge strength in the toe of Brownhill. Methods from which results can be drawn are
presented (some cannot be used due to model construction, see disadvantages within Tables 7.1 and 7.2).

Limiting conditions:
Situation Case Strength on rock- Strength on Analysed by:
Strength on SF
bridge OHWF
cohesion = 0
1 6 Purely cohesive - All methods
friction = 20°
Cohesion and Jacob models
cohesion = 0
2 9 friction - Phase2 model
friction = 20°
(friction = 40°) 1*
cohesion = 0 cohesion = 150 kPa Jacob model 1
3 4 Purely cohesive
friction = 20° friction = 20° Phase2 models
Cohesion and Jacob model 1
cohesion = 0 cohesion = 150 kPa
4 5 friction Phase2 model
friction = 20° friction = 20°
(friction = 40°) 1*

*Phase2 Model 2 could have been used in these cases. Considerably more time would

independently determine the c of the rock-bridge whilst fixing the .


have to be spent to find the limit point as the automated SSR cannot be used to

180
CHAPTER 7: Case study slope 5 (Kalgoorlie)

All models, excluding Jacob Model 1, are able to and give similar results for the
cohesive strength of rock bridge (ranging from 300 to 500kPa) for situation 1.

As Figure 7.9 illustrates, it appears the Jacob Model 1 (using slices/Janbu method –
see Appendix C), gives inappropriate values for cohesive strength in situations 1 and 2;
this is likely to be related to the high horizontal forces caused by the final three slices
located above the SF. However when some shear strength was applied to the SF,
Jacob Model 1 gave a result that was comparable to that from Phase2 modelling.

Figure 7.9: Summary of results*, from the different techniques used in the back-analysis of the
potential cohesive strength for the rock-bridge at the toe of Brownhill.

*Detail concerning the respective strengths assumed within each situation is presented
in Table 7.12, along with methods used to analyse the strength.

The rock-bridge strength has been mainly analysed with regards to potential cohesive
strength as it could be suggested that cohesive strength of the rock-bridge is likely to
be more important in providing resistance than frictional strength. This is due to the
failure of a relatively small mass with limited confining stress. Despite this, analyses
have been conducted to consider the influence of frictional strength on the back-
analysed component of cohesive strength of the rock-bridge using the Jacob Models
and Phase2 Model 1 are able to consider the influence of the rock mass  on the back-
analysed cohesive strength rock-bridge. As presented in Figure 7.9, these showed that
frictional strength upon the rock-bridge, can in fact have a substantial influence,
depending on the shear strength applied to the SF:
181
CHAPTER 7: Case study slope 5 (Kalgoorlie)

 The cohesive strength of rock-bridge can range from 110 to 485kPa when the
rock-bridge is purely cohesive;
 When 40° is applied to the rock-bridge, the cohesive strength is considerably
lower (0 to 170kPa).

The detail, from which these ranges are derived, is presented in Table 7.13.

Importantly Jacob Model 1 and Phase2 Model 1 were used to analyse the strength of
the rock bridge when a degree of friction is applied and there is the minimum strength
applied to SF (situation 4). Both models demonstrated that failure is unlikely;
consequently the minimum strength conditions suggested by Hewson (2008, personal
communication), are not plausible unless there is zero frictional strength within the
mass comprising the rock-bridge (as in situation 3).

Alternatively, discontinuity strengths within the ranges suggested are more realistic if
there is another condition, out of those suggested in Section 7.4.2, which is significantly
increasing instability.

Table 7.13: Comparison of results from selected methods of analysis, which can consider the influence of
frictional strength, on the rock-bridge within the toe of Brownhill.

Frictional

to rock-bridge 
strength applied 0° 40°

Other conditions:
analysis 
Method of Back-analysed cohesive
Strength applied Strength
strength of rock-bridge (kPa)
to OHWF applied to SF
Jacob Model 2 300 170
cohesion = 0 cohesion = 0
Phase2 Model 1 485 65 friction = 20° friction = 0°
Phase2 Model 2 444 *
Jacob Model 2 110 0
cohesion =
2 cohesion = 0
Phase Model 1 165 0 150kPa
friction = 20°
friction = 20°
Phase2 Model 2 111 *

*Phase2 Model 2 could have been used in these cases however considerably more

used to independently determine the c of the rock-bridge whilst fixing the .


time would have to be spent to find the limit point as the automated SSR cannot be

182
CHAPTER 7: Case study slope 5 (Kalgoorlie)

7.4.2 Conclusions
The failure mechanism for Brownhill is assumed to have involved fracturing through a
6m rock-bridge causing toe-breakout and subsequent sliding on the OHWF, with
kinematic release of the failure block via the Subvertical Fault. Firstly, various limit
equilibrium approaches were used; importantly the slice-based solution used had to be
modified to achieve results that compared well with the other analyses. Finally, further
analysis was through a relatively simple level of numerical modelling, which predicted a
slightly higher level of cohesive strength within the rock-bridge.

Importantly all methods could be used to derive the strength of the rock-bridge when it
is purely cohesive, and there is no strength applied to Subvertical Fault (situation 1 in
Section 7.4.1). Under these conditions, the analyses showed that the cohesive
strength of rock-bridge can range from 300 to 500kPa. Strength was lower (0 to
170kPa) when a frictional angle of 40° was added to the rock-bridge.

The ranges in the back-analysed strength highlighted that, as recommended, there


needs to be a finer definition on discontinuity strength and more detail on the structure
within the toe region. This may reveal additional issues that may affect stability, and
therefore change the back-analysed cohesive strength for the toe-breakout region.
Consequently the difference may decrease between the high mass strength (1600kPa),
suggested for a blast-damaged Peringa Basalt by PSM (2004), and that analysed
during this research ( 500kPa). Notably Pells Sullivan Meynink Pty Ltd. used RocLab
(Rocscience, 2008) to derive mass strength properties from intact strength, given a
disturbance factor of 1. As discussed in Section 2.5.1, RocLab is based on empirical
relations, via a qualitative assessment of GSI alongside the Hoek-Brown failure
criterion. Therefore it could be argued that the back-analysis approach via modelling
that is presented in this investigation, could potentially give a more reliable
representation of mass strength of blast damaged Peringa Basalt.

During the visit to the Fimiston Open Pit it was clear that micro-structure (foliation)
within this rock-bridge could have influenced the mass properties and possibly
promoted fracturing. Compared to the compressive strength of intact basalt, the back-
analysed cohesive strength is low. As outlined above this implies that the rock mass in
the toe-region is blast-damaged. This can be appreciated when considering the
influence of the damage parameter (D) in RocLab. As in Equations 2.10 and 2.11, the
D-factor affects both the mb and the s value, which decreases the cohesive strength to
a value that is more like those which have been revealed from back-analysis.

183
CHAPTER 7: Case study slope 5 (Kalgoorlie)

The analyses that have been conducted could not incorporate micro-structure, with the
ability to consider only a localised equivalent continuum mass strength for the rock-
bridge. It is a result of this, which may have caused the cohesive strength of the rock-
bridge to be so low, between zero and 500kPa (depending on the frictional conditions
on the rock-bridge and the strength applied to the OHWF and the SF).

Further evaluation of shear strength of the OHWF and SF is necessary if the back-
analysed cohesive strength of the rock-bridge is to be used in further analysis.
Importantly given the strength advised by KCGM for the SF, failure is only possible if
the strength is at the minimum of the range of strengths supplied. Some analyses have
shown that a higher cohesive strength for the rock-bridge can exist only if there is zero
strength applied to the SF.

In addition, the major underlying fault, illustrated in Figure 7.1, has the potential to
place an increased degree of mobility in the overall slope and therefore increase the
instability. Therefore it is suggested that further analysis should investigate: if the
cohesive strength of the rock-bridge can be significantly higher given this additional
potential instability; and also re-examine the influence of the respective shear strengths
of the OHWF and SF given the increased degree of instability from the underlying fault.

All models have only been conducted in two-dimensions; from above the failure looks
like it has the potential to be a wedge, therefore a model in three-dimensions could be
more beneficial. For this, one would need a more detailed geometry, which could be
sourced from remote mapping (laser scanning or photogrammetry).

Therefore the recommendation is to take analysis further, developing possibly a more


accurate model, via increasing understanding in the following areas:

 More detailed structural model, incorporating micro-structure especially within


the toe region
 Increased testing on samples from the OHWF and the SF, allowing increased
confidence on discontinuity strength
 Testing on samples from the major underlying fault, to give a range of strength
with which a sensitivity analysis can be conducted
 Re-definition of failure geometry, allowing potential modelling of a three-
dimensional structural model of the failure zone.

184
CHAPTER 8: Case study slope 6 (large slope modelling)

CHAPTER 8 – Case Study 6: Modelling of a


discontinuous mass-controlled failure at large slope
scale

8.1 Introduction
The FracMan-ELFEN approach has been successfully applied to the assessment of
pillar strength (Elmo, 2006; Pine et al., 2006; Flynn and Pine, 2007). To take this
concept further, this Chapter uses the approach to represent the mass strength,
providing equivalent continuum properties that are used to model large slope
behaviour. Detail on the FracMan-ELFEN sequence, with ways in which mass strength
has been represented and subsequently evaluated using different methods of slope
stability analysis, is presented within Figure 8.1. Comparisons are made between the
FracMan-ELFEN approach and an empirical-based approach; Figure 8.1 shows at
which stage these comparisons can be made.

The primary step of this process requires precursor characterisation of a rock mass;
the Hoptonwood Limestone mapped within Middleton Mine in Derbyshire, forms a
suitable case study, considering the previous mapping and FracMan modelling
conducted by Elmo (2006). ELFEN is used to simulate large-scale biaxial tests
numerically, subsequently deriving equivalent continuum properties to represent the
mass strength of the Middleton DFN. Accompanying the FracMan-ELFEN (numerical)
approach, an empirical method to arrive at rock mass strength is also detailed within
this Chapter, to allow comparison with the numerically derived rock mass strength.

Section 8.4 summarises a slope design process, where different methods are used to
model the equivalent continuum. The different approaches for rock slope assessment
are summarised within Section 8.5, together with an evaluation of the FracMan-ELFEN
approach for slope evaluations.

185
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.1: Sequence within the FracMan-ELFEN (numerical) approach, used to derive and analyse
rock mass strength in a large slope; an empirical-based method is also presented for comparison. (There
2
was insufficient time to analyse an empirical-based Phase model, consequently this is bracketed).

8.2 Formation of a case study slope


Using ELFEN one can perform large scale two-dimensional biaxial tests on fractured
rock with the aim to derive a compressive strength for the rock mass. Consequently
multiple biaxial tests can be used to generate a strength envelope for an equivalent
continuum, which ultimately represents the discontinuous nature of the rock mass.

As discussed in Section 2.5.1, conventional chart-type rock mass classification


approaches consider the rock mass as a single homogeneous mass with the intact
strength downgraded to represent the mass strength based on empirical data. The
proposed FracMan-ELFEN approach aims to derive the rock mass strength via direct
186
CHAPTER 8: Case study slope 6 (large slope modelling)

numerical modelling of a discontinuous mass; consequently comparisons will be drawn


between the two methods.

When analysing a large-scale fractured rock slope as an equivalent continuum, it is


important to consider the dominant orientation of discontinuities. A weak mass
strength is derived from biaxial tests where the dominant discontinuity orientation
parallels the direction of maximum principal stress (σ1). This can either occur within the
upper part or toe of a fractured slope where there is a dominant fracture orientation.
This is because respective biaxial sections have to rotate with depth in order to
simulate suitable mass shear strengths (as illustrated later in Figure 8.2).

The rock mass selected for the large-scale slope stability analysis using ELFEN, was
the Hoptonwood Limestone. As discussed in Section 8.2.1, Elmo (2006) mapped the
fracture network finding four sets of discontinuities:

 Two dominant orthogonal (north east and north west striking) sub-vertical sets;
 a third set dipping at 45° north and south; and,
 finally a set of widely spaced (between 2 and 4m) sub-horizontal bedding planes.

Therefore there are few horizontal fractures and numerous vertical and subvertical
fractures, providing a dominant fracture orientation. This means that the weakest
region of the Hoptonwood Limestone is within the upper part of equivalent continuum
slope model, where the dominant shear stresses occur approximately 30° from σ1, in
alignment with the likely mass failure surface.

This orientation alone has been used to derive equivalent continuum based properties
for the Hoptonwood Limestone, resulting in a conservative design due to weak mass
strength. In addition this investigation has also reviewed slope mass strength within a
‘zoned’ model, where the higher strength in the toe of the slope is included. As
illustrated in Figure 8.2, rotation of the biaxial stresses σ1 and σ3 relative to the same
failure pattern was necessary within the toe of the slope.

The relative rotation of biaxial stresses to the fracture orientation within the
Hoptonwood Limestone DFN results in the lower part of the failure surface being
stronger, due to shear against the dominant fracture orientation. The shear strength in
the middle part of the failure surface would be intermediate between upper (weak) and
lower (strong) values for the Hoptonwood Limestone. The strength for just the weakest
and strongest zones is derived within Section 8.3; consequently they are considered
within a slope model, as described in Section 8.4.1.

187
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.2: Illustration of how large-scale ELFEN biaxial models (incorporating a DFN) are
used to derive slope mass strength represented by an equivalent continuum. Zoning of
mass strength is included, to allow the analysis of changing shear strength with depth.

8.2.1 Locality of previous mapping and basis of model


The DFN for the Hoptonwood Limestone created by Elmo (2006) is based on field
mapping of pillars at the Middleton Limestone Mine, Derbyshire, which was operated
by OMYA. The mining method was room and pillar with large spans (17m2 pillars and
13m wide rooms), as shown in Plate 8.1. This allowed the consideration of fractures at
large scale and development of an appropriate DFN, as shown in Figure 8.3.

Plate 8.1: Typical view of Middleton Limestone Mine in Derbyshire (after Elmo, 2006).

188
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.3: DFN for a pillar in Middleton Limestone Mine, with presentation of the derived jointing on
pillar faces (after Elmo et al., 2005).

8.2.2 Previous uniaxial and biaxial ELFEN Modelling


Elmo (2006) selected seven different 2.8 x 7m sections (RA to RE, as well as P13B
and P13C), from the pillar faces presented within Figure 8.3; this gave the results in
Table 8.1. In addition to these uniaxial models, Elmo (2006) also conducted some
biaxial tests, but only on sections RA and RB. From the biaxial tests Elmo (2006)
derived two mass strengths that closely agreed with the limits suggested by RocLab
(Rocscience, 2008):

1. A mi of 12 and GSI between 40-50, matched the envelope from biaxial testing
of pillar model RA.
2. A mi of 12 and GSI between 70-80, matched the envelope from biaxial testing
of pillar model RB.

Throughout the uniaxial and biaxial testing within this thesis the same modelling
parameters, model setup and stress monitoring points were used that were originally
calibrated by Elmo (2006). However the rock mass strength presented within this
Chapter, is derived from a slightly higher axial strain than Elmo (2006) used. Elmo
(2006) completed mostly uniaxial tests to interrogate pillar strength, with a few biaxial
tests to check the FracMan-ELFEN results. The biaxial compressive strength that
Elmo (2006) determined, was obtained at relatively low axial strains (<1%). It could be
suggested that a mass strength taken from low strain is more appropriate for the
interrogation of pillar strength, where low strains can cause significant damage.

For the purpose of this investigation, higher strains have been investigated.
Consequently the biaxial results presented by Elmo (2006) can only be used as a
guide. Importantly, a mass strength has been derived for a horizontal section through

189
CHAPTER 8: Case study slope 6 (large slope modelling)

the DFN of the Hoptonwood Limestone, to allow for consideration of change in shear
strength with depth, as discussed previously.

Table 8.1: Summary, collated by Z. Flynn (2007, personal communication), of results from the simulations
of Middleton Mine pillars presented in Elmo (2006). (Further details are included in Section 8.3.1).

RA RB RC

P21=2.62 m/m2 pk= 3.27 MPa P21=1.80 m/m2 pk= 11.03 MPa P21=2.60 m/m2 pk= 2.91 MPa
Initial Final Initial Final Initial Final

23 fractures pk = 0.42 % 14 fractures pk = 0.75 % 19 fractures pk = 0.77 %

RD RE P13B

P21=2.63 m/m2 pk = 2.94 MPa P21=2.66 m/m2 pk = 4.40 MPa P21=2.95 m/m2 pk = 0.65 MPa
Initial Final Initial Final Initial Final

23 fractures pk = 0.77 % 18 fractures pk = 0.75 % 24 fractures pk = 0.53 %

P13C Using Mohr-Rankine


Models P13B and P13C are orthogonal to RA-RE.
P21=2.89 m/m2 pk = 3.65 MPa
Initial Final

23 fractures pk = 0.75 %


Run by D.Elmo version 3.7.0, pictures from Elmo (2006).

190
CHAPTER 8: Case study slope 6 (large slope modelling)

8.3 Analysis of mass strength


The following section considers the implications of intact strength and the DFN, on the
mass strength of the Hoptonwood Limestone, using both numerical and empirical
techniques. For the numerical method it is necessary to conduct a range of biaxial
tests on models that incorporate the Middleton intact rock strength and DFN. What
follows is a description of the process by which the biaxial models were created,
followed by the results from the FracMan-ELFEN approach. Finally there is a section
outlining the mass strength from an alternative empirical-based method.

8.3.1 ELFEN Biaxial Tests


Three different sections containing the Middleton DFN have been simulated during the
ELFEN biaxial testing. Two models (RC and RE) were selected from the range of
simulations that were originally conducted by Elmo (2006), as were shown in the
previous section within Table 8.1. RC and RE were chosen for further analysis as they
were representative of the middle of the strength range that Elmo (2006) reported from
ELFEN uniaxial modelling. The final section, referred to as the horizontal model, was
created from a horizontal section through one of the large (14 x 7m) Middleton pillar
models, in order to allow consideration of biaxial stress orientation and consequent
changing shear strength along the failure surface, as introduced previously within
Section 8.2.

Model Composition
All parameters and loading rates were kept at the same values that were used by Elmo
(2006), as shown in Table 8.2; these properties were applied consistently throughout
all of the biaxial modelling. Also the models were run using ELFEN analysis version
3.3.31, as it was found that using the then current version of analysis (3.9.1)
significantly, different results were attained, as shown in Appendix J.1. The only way to
accurately use the current version of ELFEN would be to re-calibrate the model
parameters (loading function, damping and normal penalty) within the new version of
analysis (3.9.1), in the same way that Elmo (2006) did for version 3.3.31.

191
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.2: Properties for ELFEN modelling parameters, derived from the uniaxial
simulations carried out by Elmo (2006).

Parameter Value
Young’s Modulus 20 GPa
Intact-material properties
Main
Poisson's Ratio 0.23
properties
Density 2600 kg/m3
Cohesion 9 MPa
Mohr- Friction angle 40°
Coulomb +
Dilation angle 5°
Rotating
Crack* Tensile strength 3.84 MPa
Fracture energy 19.47 J/m2
Contact damping 30 %
Discontinuity and mesh-

Field 0.04 m
based properties

Normal penalty (Pn) 1 GPa/m


Discrete Tangential penalty (Pt) 0.2 GPa/m
global
properties Zone 0.2 m

Smallest element 0.2 m


Friction 31°
Cohesion 100 kPa
Unstructured, Mesh density at boundaries 0.2 m
Delaunay,
details
Mesh

linear, Mesh density on fractures 0.2 m


triangular,
standard Mesh spheres 0
mesh
Stage 1 Gravity load (following sine-curve) 0 to 0.5 s
sequence
Model

σ1 loading (at 0.93 cm/s) From 0.5 s


Stage 2 σ3 loading (following respective
From 0.5 s
average σ1 response)
During stage 1 10%
Damping constraint
During stage 2 0%

*The constitutive model chosen to represent the material is the ‘Mohr-Coulomb with
tensile Rankine cut-off,’ as discussed in Appendix B.

192
CHAPTER 8: Case study slope 6 (large slope modelling)

A fundamental issue, which was discovered during the ELFEN analyses, was that the
restart (pause) function could not be used within ELFEN analysis version 3.3.31. Using
this version of ELFEN, an artificial peak in stress is generated at the beginning of a
restarted simulation, as briefly discussed within Appendix J.2. The consequence of this
was that all the models which had been restarted had to be re-run, ensuring that there
was no interruption of the simulations. This gave accurate results for the ELFEN
biaxial models, which are presented within Appendix J.3 and summarised in Section
8.3.2.

Elmo (2006) loaded the biaxial models using an applied displacement (1), with the
both platens closing by 0.07m over 7.5s; therefore 1 loading was at a rate of
0.0093m/s at each platen. The loading of the constraint (σ3) was performed as a ‘face
load’ (which is equivalent to a pressure), which follows a loading rate that is
independently matched to the average axial stress caused by the 1 loading in the
respective uniaxial model. Consequently the σ3 load rate is different for each model,
depending on how the individual section responds the 1 loading.

As shown by the stress-strain plot in Table 8.1, Elmo (2006) did not need to run
simulations for very long, as uniaxial failure generally occurred between 0.4% and 1%
strain. Consequently, simulations were set to run to a maximum of 8s runtime (which
equates to 2% axial strain). At first this was carried forward into the confined
simulations of this research; however as presented in Appendix J.3, a larger value of
maximum strain was needed to detect failure for highly confined models. Therefore,
following the primary simulations (in this case the RC model), it was ensured that the
other models (RE and horizontal models) were run at higher levels of axial strain.

During all of the biaxial tests conducted in ELFEN, the vertical stress was tracked at
several points running along the axis of the model; this followed the convention that
Elmo (2006) had set up. The calculation of shear strength was taken from an average
of the axial stress over several points. The mesh and axis along which the monitoring
points were located, is illustrated in Figure 8.4.

193
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.4: Typical mesh (with fractures embedded), used in


ELFEN biaxial simulations. Annotation shows the axis along
which stress is monitored, over several points, and
subsequently averaged.

Horizontal Models
A horizontal section was taken through the FracMan Middleton model, originally
created by Elmo (2006), and rotated as shown in Figure 8.5. This considered the
concept that the critical shear direction can change with depth, in a large slope with a
predominant vertical fracture pattern, (as was previously shown in Figure 8.2).

Figure 8.5: Large (14 x 7m) Middleton mine pillar model from Elmo (2006), from
which a horizontal section is extracted and rotated to create an ELFEN biaxial
model with a predominantly horizontal fracture pattern.

A second horizontal model was created upon finding that the section presented in
Figure 8.5, behaved in a surprisingly weak manner. This was due to a number of
adversely orientated discontinuities that were controlling the failure mechanism, as
illustrated in Figure 8.6a, where a failed block is highlighted and other potential weak
areas are indicated. Subsequently, in order to consider the behaviour of a presumed

194
CHAPTER 8: Case study slope 6 (large slope modelling)

stronger model, the discontinuities highlighted in Figure 8.6b, were removed from the
specimen and the model was re-run to give the results that are presented in the latter
part of the following section.

Figure 8.6: Fractured state of horizontal section when model is run with
default fractures (a); fractures highlighted in (b) are removed to allow the
predominant fracture orientation to control failure.

8.3.2 Results from the FracMan-ELFEN Approach


A peak in the average axial stress of each biaxial test, gives a value for the strength of
the mass under the respective confinement conditions. As a result of the fracture
network, the classic peak strength behaviour that occurs within intact uniaxial tests
does not occur within tests on fractured rock. Consequently in some simulations it is
difficult to pick the actual yield points, with peaks occurring over a wide range of axial
strains depending on the confining pressure used.

During the biaxial simulations within ELFEN, it was noted that many of the models had
two peaks. The peak that occurs at the lower axial strain indicates the onset of failure.
The final actual peak in the maximum principal stress (σ1max) is exposed at higher axial
strain. Consequently it was decided that two data sets would be collected from each
model, one reflecting the primary peak in σ1 that occurs at axial strains of up to 1.5%,
and the other from the σ1max that occurs at axial strains of up to 2.5%.

The pre-peak strength derived from axial strains of up to 1.5%, denotes a situation that
is controlled mostly by the Young’s modulus, falling on the pre-peak (stiffness) stress
strain curve. The data set from the σ1max, which occurs at axial strains of up to 2.5%, is
more likely to reflect the elastic-plastic post-yield point of each of the models. Note that
in some biaxial models only one data set can be retrieved, this is particularly the case
within the unconfined runs, where the yield point of the mass is achieved quickly (at low

195
CHAPTER 8: Case study slope 6 (large slope modelling)

degrees of axial strain). Consequently where this occurs, the single σ1max strength
value is used in both the 1.5% and 2.5% axial strain failure envelopes.

Confinement has a significant influence on the model, with large differences between
the low strength unconfined models, and the very strong highly confined simulations. It
has been found that this can cause skewing in the data analysis, with negative values
for the intercept (and consequent Hoek-Brown s value). In these cases the Hoek-
Brown envelope has been forced through zero, causing s to be zero, as reflected in
Table 8.3.

Primary simulations on all of the sections have shown that the Hoptonwood Limestone
has a high mass strength. The data from these simulations is presented in Appendix
J.3, and a summary of the results is presented in Figure 8.7.

Figure 8.7: (a) Presents data that marks the onset of failure in the Hoptonwood
Limestone, derived from peaks in σ1 which occur at low strain. (b) Presents the yield
behaviour of the respective sections, which occur at higher degrees of axial strain.

196
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.3 presents the range of mass strengths, which can be derived from data
presented within Figure 8.7. Note that the respective lines of best fit and calculus,
which were used to derive the strength properties listed in Table 8.3, are presented in
Appendix J.3.

Table 8.3: Hoek-Brown and Mohr-Coulomb properties from non-downgraded RC, RE and horizontal
ELFEN biaxial models.

Derived Mohr-
Derived Hoek-Brown
Mass strength taken from Coulomb properties
constants from
Model peaks in vertical stress that ELFEN biaxial tests:
from ELFEN biaxial

 (°)
occur: tests:
s mb c (MPa)
up to 1.5% axial strain 0.0163 7.286 1.1 57
RC
up to 2.5% axial strain 0 9.472 0.9 59
up to 1.5% axial strain 0.0274 2.138 1.4 46
RE
up to 2.5% axial strain 0.0974 5.812 1.8 55
up to 1.5% axial strain 0.0704 6.881 1.8 55
Horizontal
up to 2.5% axial strain 0 * 1.1 63

* This result gave a very high mb due to the high strength at high confinement, as
presented in Appendix J.3. It is considered anomalous and therefore not included
within the Table.

As can be seen from Table 8.3, the results from the RE and RC simulations, give a
considerable range in the m value for the mass (mb), c and , depending on the degree
of axial strain at which the strengths are taken. The database in RocLab (Rocscience,
2008), suggests that the Hoek-Brown constant mi for a crystalline limestone should
be 12  3.

The use of the non-linear Hoek-Brown criterion, to represent the equivalent continuum
derived from the FracMan-ELFEN approach, is discouraged by the variability in the mb
parameter and the anomalous result when there is high strength at high confinement.
Alternatively it is suggested that a linear approach is more appropriate, such as the
Mohr-Coulomb criterion. This is explored further below, and discussed within the
synthesis at the end of this section.

Downgraded (intact strength) models


The biaxial models reported in the section above have revealed mass strengths that
are too high to allow circular failure through the rock mass of a 1000m slope, as was
found when the weakest case was modelled in SLIDE (Rocscience, 2008). This is
realistic considering the intact strength of Hoptonwood Limestone. This could indicate
that failures within a rock mass of this nature are likely to be structurally-controlled

197
CHAPTER 8: Case study slope 6 (large slope modelling)

(involving step-path/progressive failure, as discussed in Section 2.2.5), rather that


mass-controlled.

To allow the further application of the FracMan-ELFEN process and assess a rock
mass in which the DFN has been explicitly modelled, the intact strength of the
Hoptonwood Limestone was degraded. This gave a new rock mass, which retained
the original DFN created for Middleton Limestone mine by Elmo (2006). Large-scale
mass-controlled (circular) failure can be considered within this new downgraded rock
mass, allowing further application of the equivalent-continuum (FracMan-ELFEN)
approach.

To obtain the downgraded rock mass, the Mohr-Coulomb and Rotating Crack intact
strength properties of the original models were quartered to give the new ‘downgraded’
intact strength, as presented in Table 8.4. Note that the elastic properties (Young’s
Modulus, Poisson’s Ratio and the density), have all remained at the same values for
the downgraded models, as they were within the original models, (listed in Table 8.2).

Table 8.4: Original (from Elmo, 2006), and quartered intact strength properties for ELFEN biaxial
models with the later being for the weaker/downgraded models.

Strength values for biaxial


Parameter models:
Original Downgraded
Cohesion 9 MPa 2.25 MPa
Intact-material

Mohr-
properties

Friction Angle 40° 11.8°


Coulomb
+ Dilation Angle 5° 5°
Rotating Tensile Strength 3.84 MPa 0.96 MPa
Crack
2
Fracture Energy 19.47 J/m 4.86 J/m2

Later in this Chapter the Hoek-Brown mass strength derived from the FracMan-ELFEN
approach, is used in a limit equilibrium method (SLIDE). This requires consideration of
the unconfined compressive strength (UCS); in the case of the Hoptonwood Limestone
the UCS was 48MPa, as reported later in Table 8.8. Because a downgraded intact
strength was analysed, this also had to be quartered, giving a UCS of 12MPa.

Both RE and horizontal models were re-analysed using the new downgraded intact
strength, giving the stress-strain results presented in Appendix J.4; the summary of
these biaxial tests is given in Figures 8.9, with the resulting derived mass strength
listed in Table 8.5. This new (downgraded) mass strength is now low enough to enable

198
CHAPTER 8: Case study slope 6 (large slope modelling)

the consideration of circular failure through the Hoptonwood Limestone DFN as


outlined in Section 8.4.

Figure 8.8: (a) Presents data that marks the onset of failure within downgraded
material, derived from peaks in σ1 which occur at low strain. (b) Presents the yield
behaviour of the respective sections, which occur at higher degrees of axial strain.

Table 8.5: Hoek-Brown and Mohr-Coulomb properties from RE and horizontal ELFEN biaxial models on
the downgraded Hoptonwood Limestone.

Derived Mohr-
Derived Hoek-Brown
Mass strength taken from Coulomb properties
constants from
Model peaks in vertical stress that ELFEN biaxial tests:
from ELFEN biaxial

 (°)
occur: tests:
s mb c (MPa)
up to 1.5% axial strain 0 2.553 0.8 48
RE
up to 2.5% axial strain 0.0004 4.180 0.9 53
up to 1.5% axial strain 0 11.969 0.8 62
Horizontal
up to 2.5% axial strain 0.0130 * 1.1 64

* This result gave a very high mb due to the high strength at high confinement, as
presented in Appendix J.4. It is considered anomalous and therefore not included
within the table.
199
CHAPTER 8: Case study slope 6 (large slope modelling)

In addition to the data presented within Table 8.5, a further higher confinement (4MPa)
simulation was conducted, although only in the RE section. This was intended to
increase understanding of mass-strength within the downgraded Hoptonwood
Limestone. However, this simulation alone took 15 days to run, with an actual σ1max
occurring at around 6% axial strain, as presented in Figure 8.9.

Figure 8.9: Stress-strain plot of downgraded RE biaxial models, with data from a 4MPa confined
simulation; screen shots illustrate the state of fracture at which the strengths for the respective
models were taken.

Importantly by running a model confined by 4MPa, it becomes clear that the stress-
strain response of the mass is linear and therefore produces a non-linear relationship
upon a Hoek-Brown plot, as illustrated in Figure 8.10a. This can be used to suggest
that the Mohr-Coulomb strength criterion is a more appropriate approach to assessing
mass strength in this case.

200
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.10: (a) Data from downgraded RE model, demonstrating an exponential


envelope upon a Hoek-Brown plot; consequently a Mohr-Coulomb failure envelope is
more appropriate as presented in (b).

As it has been shown that a Mohr-Coulomb strength criterion is more appropriate, the
data for the downgraded models has been re-interpreted to derive a range of Mohr-
Coulomb strength parameters; these are presented in Table 8.6. Different Mohr-
coulomb parameters can be taken from the respective envelopes depending on
whether the envelope is created from all or a selection of the confined models. Figure
J.7 in Appendix J.4 gives detail on the failure envelopes and subsequent calculus
involved, in deriving the mass strength parameters presented in Table 8.6.

Table 8.6 demonstrates that there is a range in the derived c and  of the mass,
depending on the range of σ3 chosen to form the respective Mohr-Coulomb failure
envelope.

201
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.6: Mohr-Coulomb properties from downgraded RE, and downgraded Horizontal, FracMan-
ELFEN biaxial models.

Orientation of dominant Data taken from following Resulting Mohr-Coulomb Strength


range of σ3 to create
Friction ()
discontinuities within parameters:
downgraded Model: envelope (MPa): Cohesion (kPa)
0, 0.5, 1, 2 and 4. 410 58
0, 1, 2 and 4. 164 58
Vertical (RE)
0, 0.5, 1 and 2. 876 53
0, 1 and 2. 577 54
0, 0.5, 1 and 2. 1083 64
Horizontal
0, 1 and 2. 546 66

Hoek & Bray Empirical Chart Analysis


The Mohr-Coulomb mass strengths derived from the ELFEN biaxial testing presented
in Table 8.6 were used to determine the FOS for a slope of certain height and
inclination. To briefly review the mass strength predicted by the FracMan-ELFEN
approach, the steepest and largest slope was analysed (1000m 55° slope).

The simplest method of analysis for circular failure is to use the charts published in
Hoek and Bray (1981). The strength inputs, calculation and subsequent results from
the Hoek and Bray charts are presented in Table 8.7.

Table 8.7: Calculation and consequent FOS from the Hoek and Bray circular failure charts (1981) on a
dry 55° 1000m dry slope; note that in this case F refers to factor of safety, and H to slope height.

Data taken from Resulting Value from Value from FOS from FOS from

tan  tan 
following range Mohr- x-axis y-axis x-axis y-axis

x y F F
of σ3 to create Coulomb
HF Hx
c c
envelope Strength * *
(MPa): parameters: F y

 = 58°
0, 0.5, 1, 2 and c = 410 kPa
0.011 1.11 1.4 1.4
4.

 = 58°
c = 164 kPa
0, 1, 2 and 4. 0.005 1.27 1.3 1.3

 = 53°
c = 876 kPa
0, 0.5, 1 and 2. 0.024 0.93 1.4 1.4

 = 54°
c = 577 kPa
0, 1 and 2. 0.016 1.00 1.4 1.4

*  (unit weight) of Hoptonwood Limestone = 26kN/m3

It can be seen that for a large steep slope, the FOS values do not vary much between
cases as the shear strength is determined by the (similar) friction component. Also
where the friction angles were lower this was compensated by higher cohesion values.

202
CHAPTER 8: Case study slope 6 (large slope modelling)

Synthesis
The ELFEN biaxial testing has shown that slight differences in the fracture pattern can
have a significant influence on the strength at high confinements. An example of this is
provided in Figures 8.12a and 8.12c, where there is quite a significant difference
between the RC and RE Hoek-Brown envelopes despite the closeness of their P21
values (2.60 and 2.66m/m2 for RC and RE respectively). This is further illustrated in
Figure 8.12 by the coefficient of determination (R2), demonstrating that there is less
variability when the data is taken from high axial strain.

The same can be found when representing the data via the Mohr-Coulomb criterion,
with slightly less scatter when data is derived from higher axial strain. This
demonstrates that it is more reliable to take the strengths for biaxial models from peaks
in the maximum principal stress that occur at degrees of up to 2.5% axial strain.

To minimise variation and allow a more reliable mass strength from the FracMan-
ELFEN approach, many simulations should be run (perhaps adopting a Monte Carlo
approach) of slightly different fracture patterns (all with a similar P21). This would
provide a data set through which an average envelope could be drawn obtaining a
more reliable mass strength. In addition, simulations should be run up to 3% axial
strain, to minimise variation and reflect the peak behaviour considered appropriate for
relatively unconfined (slope) situations.

From the data for the RE model presented in Figure 8.10 it was concluded that the
Mohr-Coulomb failure envelope fitted the FracMan-ELFEN biaxial data better than a
Hoek-Brown envelope.

A way in which the FracMan-ELFEN rock mass data can be compared against an
empirical approach is to use RocLab (Rocscience, 2008). Both Hoek-Brown and Mohr-
Coulomb mass strength can be derived using this method, although as shown in
Figure 2.9, the Mohr-Coulomb mass strength obtained is heavily dependent upon the
σ3max. Therefore the following section uses RocLab to derive only Hoek-Brown mass
strength; this is compared to the FracMan-ELFEN approach despite, as previously
discussed, the limited reliability of the numerical-based Hoek-Brown mass strength (in
this case).

203
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.11: Synthesis


of all data derived from
ELFEN biaxial testing.
(a) and (b) Present the
failure envelopes from
low axial strain; (c) and
(d) indicate the results
when the mass
strength is taken from
a higher axial strain.
204

204
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.12: Data from ELFEN testing on non-downgraded RC and


RE biaxial sections forming average Hoek-Brown envelopes and
Mohr-Coulomb envelopes respectively in (a) and (b) respectively.

8.3.3 Rock Mass strength determined from conventional approaches


RocLab (Rocscience, 2008) is the current most widely used approach for determination
of a mass strength from intact strength data combined with rock mass classification
data. It was outlined within Section 2.5.1 how the empirical-based approach within
RocLab is based upon the Generalised Hoek-Brown criterion.

This section reviews the mass strength that can be derived using RocLab; the mass
strength derived is used as a comparison against the numerical FracMan-ELFEN
approach. The intact strength data for the Hoptonwood Limestone at Middleton Mine
was presented by Pine et al. (2006) and is reproduced in Table 8.8. This can be used
in RocLab to give the mass strength data presented in Table 8.9.

205
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.8: Strength inputs for Hoptonwood Limestone in Middleton Mine, obtained from Pine et al.
(2006).

Value/range of values for Hoptonwood


Property
Limestone
Intact compressive strength (σci) 48 MPa
Geological Strength Index (GSI) 60 to 70
Hoek Brown constant for intact rock (mi) 10

Guidance within RocLab (Rocscience, 2008) suggests that the mi for a crystalline
limestone, should be 12  3. Therefore the mi suggested by Pine et al. (2006) is at the
lower end of this range; the upper end of the range will also be considered within the
analysis within this section.

Table 8.9: Range of Hoek-Brown mass strength parameters for the Hoptonwood Limestone, that
have been derived using RocLab (Rocscience, 2008), for comparison against results from the
FracMan-ELFEN analysis, listed in Table 8.10.

Hoek-Brown mass strength from RocLab


GSI mi
mb s
60 2.40 0.0117
10
70 3.43 0.0357
60 3.60 0.0117
15
70 5.14 0.0357

During this study, the GSI and the mi are considered the only variable inputs within
RocLab, which cause an alteration to the Hoek-Brown mass strength and the resulting
range of strengths, are presented within Table 8.9. Table 8.10 presents a summary of
data from the FracMan-ELFEN approach.

Table 8.10: Range of Hoek-Brown mass strength parameters for the Hoptonwood Limestone, that
have been derived using the FracMan-ELFEN approach.

Mass strength taken Hoek-Brown mass strength from FracMan-


Section from peaks in vertical ELFEN Approach
stress that occur: mb s
up to 1.5% axial strain 7.286 0.0163
RC
up to 2.5% axial strain 9.472 0
up to 1.5% axial strain 2.138 0.0274
RE
up to 2.5% axial strain 5.812 0.0974
Horizontal up to 1.5% axial strain 6.881 0.0704
RE - up to 1.5% axial strain 2.553 0
downgraded up to 2.5% axial strain 4.180 0.0004
Horizontal -
up to 1.5% axial strain 11.969 0
downgraded

206
CHAPTER 8: Case study slope 6 (large slope modelling)

From comparison of Tables 8.9 and 8.10, it is clear the FracMan-ELFEN approach
provides a far wider range of Hoek-Brown mass strengths, with dependency on the
axial strain and the modelled fracture network. RocLab consistently predicts a medium
mass strength, with variability due to the initial mi and the GSI chosen. It must be
noted though, that RocLab cannot consider the significant influence of discontinuity
orientation or the degree of strain.

The mass strength derived from RocLab is only comparable to section RE, which has
the weakest of all mass strength, suggested by the FracMan-ELFEN approach.
Therefore the FracMan-ELFEN approach indicates that the mass strength of the
Hoptonwood Limestone could be significantly greater than the mass strength predicted
by RocLab. However, as discussed throughout Section 2.5.1, it can be suggested that
it is inadequate to use an empirical approach such as RocLab, to represent a blocky
rock mass.

In summary, this investigation has shown that the mass strength derived from the
FracMan-ELFEN approach is highly dependent on a number of aspects that cannot be
considered within RocLab. The variability in Hoek-Brown properties, suggested by the
FracMan-ELFEN approach, indicates that the Hoek-Brown criterion is limited in its
ability to adequately represent numerical-derived mass strength. The following section
simulates rock slope strength based on the FracMan-ELFEN approach; subsequently
Section 8.5 reviews the RocLab-derived mass strength reported within this section, to
make a direct comparison to a numerically derived slope mass strength.

207
CHAPTER 8: Case study slope 6 (large slope modelling)

8.4 Modelling of the FracMan-ELFEN equivalent continuum


The Middleton DFN was considered via ELFEN biaxial modelling outlined in the
previous section; subsequently various mass strengths were derived each representing
an equivalent continuum of the Hoptonwood Limestone. This section reviews this
strength via the implementation of commercially available limit equilibrium and finite
element software, to analyse various slopes.

8.4.1 Slide Analysis


The Hoek-Brown or the Mohr-Coulomb strength parameters derived from the ELFEN
biaxial testing outlined in the previous section, can both be directly entered into SLIDE
(Rocscience, 2008). This allows the computation of the factor of safety (FOS) for a
given slope.

Tables 8.11 and 8.12 present the results from an analysis of the rock mass using
Hoek-Brown and Mohr-Coulomb strength parameters respectively. In all situations a
1000m slope was simulated, with different overall slope angles. As shown previously
in Figure 8.2, different mass shear strengths can be incorporated when considering the
position on the failure surface within a rock mass with a dominant fracture orientation.

As outlined in Section 8.2, the dominant fracture orientation is vertical in the


Hoptonwood Limestone; subsequently a stronger toe region would be expected. This
is incorporated in a ‘zoned’ model by using a combination of the strength from the
downgraded RE and the downgraded horizontal biaxial models. Figure 8.13 illustrates
how the mass strengths were used in a SLIDE zoned model, where two thirds of the
slope is composed of weak mass strength derived from dominantly vertically (RE)
biaxial tests. The lower third of the zoned slope comprises the stronger mass strength
from the “horizontal” biaxial tests.

208
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.13: Illustration of the composition of a ‘zoned’ SLIDE model, to allow for
the rotation of biaxial stress with depth (relative to the shear strength on a large
circular failure surface).

Note that for the purpose of analysis, a homogeneous un-zoned model was also
analysed. This is composed entirely of the weak, dominantly vertically fractured (RE),
mass strength.

The results from the SLIDE analysis using Mohr-Coulomb parameters presented in
Table 8.11, are similar to those presented in Table 8.7 from the Hoek-Bray circular
charts, which were more conservative predicting a FOS 10 to 20% lower that the
equivalent model in SLIDE. Even using the weakest Mohr-Coulomb properties and the
steepest slope, SLIDE has demonstrated that circular failure is unlikely to occur within
a slope comprised of downgraded Hoptonwood limestone, despite having a reasonable
degree of fracturing (P21 = 2.6m/m2). It is only the non-zoned 55 slope model, with
Hoek-Brown strength (taken from 1.5% axial strain), that gives a FOS of less than 1.

As is shown in Table 8.11, the selection of respective biaxial strengths can have a
significant influence on rock mass strength (in this case there was only time to
investigate the influence when using the Mohr-Coulomb criterion). The strength of the
equivalent continuum formed by selected confinements, gave different depth failure
surfaces, related to rock mass strength.

Due to the high strength at high confinement, a Mohr-Coulomb SLIDE model with the
mass strength derived from a high confinement (σ3 = 4MPa), gave a shallow circular
failure surface, as illustrated in Figure 8.14. This is because when the strong
σ3 = 4MPa result is included within the data set, the Mohr-Coulomb failure envelope is
pushed steeper and thus a higher and potentially incorrect  is derived for the mass.

209
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.14: Shallow circular failure within a SLIDE model of a 45° slope with Mohr-
Coulomb mass properties c = 164kPa and  = 58°.

The mass strength, formed from data excluding the 4MPa confinement, gave an
equivalent continuum strength with a slightly lower frictional component. Subsequently
a failure surface of reasonable depth was predicted by SLIDE, as illustrated in
Figure 8.13.

Therefore within both the Hoek-Brown SLIDE modelling (Table 8.12) and also the
Phase2 simulations (outlined in the following section), the data from the high
confinement (σ3 = 4MPa) is not incorporated. In addition the data from the low
confinement (σ3 = 0.5MPa) is also excluded, because this was also strong and
otherwise contributes to a high cohesive strength within the equivalent continuum. This
left the mid-range confinements (σ3 = 0, 1 and 2MPa), providing a strength that fell
more towards the middle of the range presented in Table 8.11.

210
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.11: Results from SLIDE modelling of Mohr-Coulomb mass strength derived from the ELFEN biaxial testing.

Mass strength taken Resulting FOS for a 1000 m slope with


Source SLIDE Model SLIDE Inputs:
 ( )
from following range overall slope angle of:
Model composition
of σ3 (MPa): c (kPa) 45° 50° 55°
0, 0.5, 1, 2 and 4. 410 58 2.1 1.8 1.6
RE 0, 1, 2 and 4. 100% strength from 164 58 1.9 1.6 1.4
downgraded 0, 0.5, 1 and 2. RE downgraded 876 53 2.0 1.8 1.6
0, 1 and 2. 577 54 1.9 1.7 1.5
Zoned
downgraded 
Top RE
Horizontal 577 54
downgraded
0, 1 and 2. 2.2 1.9 1.7
and
downgraded 
RE Bottom Horizontal
546 66
downgraded

Table 8.12: Results from SLIDE modelling of Hoek-Brown mass strength derived from the ELFEN biaxial testing.

Resulting FOS for a 1000 m slope with


Mass strength taken SLIDE Inputs:
Source SLIDE Model overall slope angle of:
from peaks in vertical
Model composition Unit weight UCS
stress that occur: s m 45° 50° 55°
(kN/m2) (MPa)
RE up to 1.5% axial strain 100% strength from 26 12 0 2.553 1.1 1.0 0.9
downgraded up to 2.5% axial strain RE downgraded 26 12 0.0004 4.180 1.3 1.2 1.1

downgraded 
Zoned Top RE
26 12 0 2.553
Horizontal
downgraded
up to 1.5% axial strain 1.2 1.1 1.0
downgraded 
and Bottom Horizontal
26 12 0 11.669
RE
downgraded
211

211
CHAPTER 8: Case study slope 6 (large slope modelling)

8.4.2 Phase2 Analysis


Phase2 has a Shear Strength Reduction (SSR) module, as discussed in Section 2.3.4;
this enables the determination of Mohr-Coulomb failure parameters which apply at a
FOS of 1 (in effect a back-analysis). To create an appropriate model using Phase2, an
elastic stress field needs to be created to simulate gravity loading within the continuum.
In order to do this a suitable k ratio should be implemented; in this case k (both in and
out-of-plane) of 0.3 was calculated using Equation 2.3.

Phase2 modelling has been completed on the un-zoned slope, composed of a weak
mass strength derived from the vertically fractured (RE) Middleton DFN and includes
the determination of stress paths failure.

During the modelling, it was found that:

 The sequence of unloading had no influence on the subsequent shear strength


reduction factor (SRF) when a default coarse mesh is used.
 The mesh density significantly influences the critical SRF calculated by Phase2.
A finer mesh resulted in a smaller critical SRF, presumably due to the ease of
strain localisation. However, in order to create such a mesh within a model, it
is necessary to increase discretization and meshed element sizes manually,
until a reasonable graded mesh is created within the area of interest.

Failure has only been analysed in a compressive stress state, as mass-controlled slope
instability within an equivalent continuum is unlikely to involve significant regions of
tensile failure. Consequently a high tensile strength was used that is one tenth of the
compressive strength (in this case downgraded Middleton is under consideration,
therefore σc = 12MPa and thus σt = 1.2MPa). In addition, it was ensured the σt was not
factored with the other strength parameters, during the SSR process. Analyses were
also conducted with zero tensile strengths; however this had minimal influence on the
ultimate critical SRF calculated by the SSR.

The results from the SSR technique within the coarse and finely meshed models are
outlined in Table 8.13; note that the fine mesh had to be manually inputted and is
consequently later presented in Figure 8.15; the coarse mesh was the default that
Phase2 selected.

212
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.13: Results from a Shear Strength Reduction (SSR) study on a 55 1000m high slope; with
resulting Shear Strength Reduction Factor and consequent mass properties which can be checked in a
SLIDE model.

Resulting Mass FOS from SLIDE


Phase2 Mass
Mesh size in Critical strength from SSR (of a Slope with
Strength Inputs:
Phase2 model
 ( )  ( )
SRF for FOS = 1 mass strength
c (kPa) c (kPa) from critical SRF)
COARSE –
default from 577 54 1.47 393 43 1.0
Phase2
FINE –
577 54 1.31 440 46 1.1
manual input

From Table 8.13, it is clear that there is a close, but not exact match between the FOS
that SLIDE predicts and the corresponding SSR which Phase2 calculated; ideally the
FOS should be 1 given the mass strength determined from the SSR model. The study
into appropriate representative mesh sizes could be an area for further analysis;
although for this thesis, the 10% difference presented by the fine mesh is considered
negligible.

Stress-path monitoring within toe region


Monitoring of the stresses has only been completed within an un-zoned finely meshed
slope, which yielded the strength/FOS results presented in Table 8.13. It was found
that the excavation sequence did not influence the ultimate SRF; however the process
of unloading would clearly affect the stress paths within a slope. The stress path data
presented within the following section is from a six-stage simulation, where the first
stage is elastic gravity loading; subsequently there are five-stages of excavation, as
depicted in Figure 8.15.

Two points in the shear zone near the base of the slope were chosen for stress
monitoring, as also illustrated in Figure 8.15:

1. Point A – just outside the boundary of failed nodes, and


2. Point B - within the boundary of failed nodes closer to the face.

Later in this Section a third location was chosen in order to track stress through to
failure:

3. Point C – within the highly strained area which marks the location of potential
circular failure plane.

213
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.15: Mesh and extent of model used in the Phase2 55° large slope simulations. Annotation is
given in the zoomed view to indicate the extent of failure when SRF = 1.31.

Figure 8.16 shows the Mohr circle stress plots for Points A and B during the excavation
of the slope. Note that once the final excavation is removed (in stage 6), the SSR
procedure starts allowing the limit point and subsequent new mass strength to be
calculated, as presented previously in Table 8.13.

The critical SRF from Phase2 within this model was 1.31; from this a new Mohr-
Coulomb envelope has been plotted, as presented in Figure 8.16. However this
envelope is inapplicable to stages 1 to 6, as the mass strength that it defines only
occurs once the SSR has reduced the strength of the mass by a factor of 1.31.

The stress circle plot demonstrates that there is a re-distribution of stress within the toe
region as the excavation sequence proceeds, and σ3 decreases as expected. The
stress varies at Point B, during unloading (decreasing σ3), more than at that Point A.
This is most probably due to the proximity of Point B to the slope face, and is
dependent on mesh density. After the final excavation has been removed (in stage 6),
the SSR process then starts within Phase2, which causes only Point B to fail.

To increase understanding of the failure within the model and the consequent
progression of stress within the base of the slope, a new model was ran with a mass
strength that it is at the limit point of failure, using the properties (derived from the SRF
of 1.31) listed within Table 8.13. The progression of stress at Point B, is clearer when
plotted on a p-q plot, as presented in Figure 8.17b; note that the respective Mohr-
214
CHAPTER 8: Case study slope 6 (large slope modelling)

Coulomb envelope has been re-plotted using an equivalent cohesion and friction within
the p-q plot. Additionally, the development of shear strain within this model has also
been analysed, and scaled stress tensors have been manually plotted, as presented
later in Figure 8.19.

Figure 8.16: Mohr-Coulomb plots of stresses at two points (Points A and


B in (a) and (b) respectively) at different stages during the excavation
and subsequent SSR of a 1000m slope composed of a mass strength
from the downgraded (vertically fractured) Middleton DFN.

215
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.17: (a) Mohr-Coulomb plot of stress at Point B during the excavation
of a 1000m slope composed of a mass strength derived from a SRF of 1.31 on
the downgraded (vertically fractured) Middleton DFN. (b) Stress path derived
from the Mohr-Coulomb plot.

As is illustrated in Figure 8.15, Point B is within the boundary of failed nodes, also
during the SSR simulation the state of stress at Point B appears to exceed failure when
the mass strength from the critical SRF is taken into consideration, as presented in
Figure 8.16. However despite this, the actual stress at Point B does not reach failure
within a model that is composed of a mass strength derived from the critical SRF
of 1.31, as illustrated in Figure 8.17. Instead, for failure one has to look to a location,
such as Point C, which is within the localised area of higher shear strain as presented

216
CHAPTER 8: Case study slope 6 (large slope modelling)

in Figure 8.18. This slight discrepancy is probably due to a different slope behaviour
and stress redistribution using the lower, factored (1.31) strengths.

Figure 8.18: (a) Mohr-Coulomb plot of stress at Point C during the excavation
of a 1000m slope composed of a mass strength derived from a SRF of 1.31 on
the downgraded (vertically fractured) Middleton DFN. (b) Stress path derived
from the Mohr-Coulomb plot.

From Figure 8.18 one can see that failure actually occurs from stage 5 at Point C;
following this there is a significant decrease in both σ1 and σ3 once the final part of the
slope is unloaded, during stage 6. This can be seen in Figure 8.19, which shows the
Phase2 model that used for the stress path monitoring in Figures 8.18 and 8.19. The

217
CHAPTER 8: Case study slope 6 (large slope modelling)

strength used in the Phase2 model, was derived from a SRF of 1.31 on the
downgraded (vertically fractured) Middleton DFN, as was presented in Table 8.13.

As shown in Figure 8.19, from stage 5 there is a significant strain just above the toe of
the slope, at the locality of Point C. In addition the strain plot at stage 5 clearly shows
two circular failure paths, one extending from the boundary of the stage 4 excavation
and the second from stage 5. However, the failed nodes that are plotted, illustrate that
these circular failures have a limited extent and do not meet the surface. Consequently
it can be suggested that signs of instability, (accelerated movement etc.), are likely to
be seen from stage 4, although full development of circular failure surface is unlikely
until stage 6.

With the stress tensors plotted in Figure 8.19b, one can appreciate the following
aspects:

 From stage 4 there is a significant decrease and rotation of both σ1 and σ3


within the areas that are failed under shear, with a clear region of significantly
decreased σ3 within the base of the slope, defining the circular failure which
develops as a result of the final stage of excavation. (A discussion is given in
Section 8.5 regarding the importance of this rotation in respect to the initial
discontinuum modelling).
 There are large stress tensors adjacent to this shear zone, illustrating that the
shear zone is well-defined.
 There is a clear rotation of stress tensors within the toe of the slope during
failure.

218
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.19: (a) Shear strain and failed nodes, during the final three stages of excavation of a 1000m slope. (b) Maximum principal stress and failed
219

nodes of the same model and stages as in (a), with scaled stress tensors to illustrate the orientation and ratio between σ1 and σ3. 219
CHAPTER 8: Case study slope 6 (large slope modelling)

8.4.3 Influence of Groundwater on Slope Strength


There are a number of ways in which the influence of groundwater can be considered;
a quick and basic (empirical-based) approach is to use circular failure charts presented
in Hoek and Bray (1981). It must be noted at this stage, that Hoek and Bray (1981)
suggest that the charts are used only as a rough guide, preceding a more detailed
analyses. Alternatively, groundwater can be incorporated into a numerical model; both
SLIDE and Phase2 (Rocscience, 2008) can simulate slopes with a phreatic surface. In
this research only SLIDE has been used. Comparisons are drawn between the
resulting FOS derived from both the Hoek and Bray charts and SLIDE, as presented in
Table 8.15. The strength inputs for both these methods are listed below, and the
calculus for the charts is presented in Table 8.14.

H (Height of slope) = 1000m


 (unit weight of Hoptonwood Limestone) = 26kN/m3
c (cohesive strength of Hoptonwood Limestone) = 577kPa
 (frictional strength of Hoptonwood Limestone) = 54°

Table 8.14: Calculus and consequent FOS from the Hoek and Bray circular failure charts (1981); note
that in this case F refers to factor of safety.

Value from Value from FOS from FOS from

tan  tan 
Hoek and Bray circular x-axis y-axis x-axis y-axis

x y F F
failure chart number

HF Hx
c c
(groundwater conditions)
F y
1 0.017 1.05 1.3 1.3
4 0.022 1.36 1.0 1.0
5 0.030 1.90 0.7 0.7

Note that for the SLIDE models of chart 3 and chart 4, the phreatic surfaces had to be
input by hand to follow a curve close to the slope face, which provides a close-as-
possible match to the illustrations of the curved phreatic surfaces, provided by Hoek
and Bray (1981).

Table 8.15: Influence of groundwater on the stability of a 55 1000m slope composed of material
derived from ELFEN biaxial modelling on vertically aligned Middleton DFN.

Mohr-Coulomb Strength Phreatic


Minimum
parameters: Surface from FOS from Slope
Friction ()
FOS from
Cohesion Hoek and Bray Stability Charts
SLIDE
(kPa) chart number
1 – dry 1.3 1.5
3 – 50 % wet 1.2 1.2
577 54
4 -75% wet 1.0 0.9
5 – 100% wet 0.7 0.5

220
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.15 demonstrates there is up to a 20% difference between the FOS resulting
from the Hoek and Bray charts and the FOS that SLIDE predicts. As illustrated in
Figure 8.20, this could be due to the different critical failure surfaces that each method
predicts for a 1000m 55° slope. The critical surfaces predicted by the Hoek and Bray
(1981) charts, have to fail through the toe of the slope, which as predicted in Figure
8.20a, is not necessarily accurate for dry conditions. Figure 8.20b shows that the Hoek
and Bray (1981) charts forecasts a larger critical plane than SLIDE, with the same FOS
as the smaller circular failure plane predicted by SLIDE. This could be due to
inconsistency between curvatures of the phreatic surface within the two methods, as
Hoek and Bray (1981) do not provide guidance on the detail of the curve for the
phreatic surface for each chart; instead this has to be estimated from the illustrations
they present.

Figure 8.20: Comparison of critical failure planes predicted by SLIDE and the Hoek and Bray (1981)
circular failure charts, when different groundwater conditions are assumed.

As previously stated, the Hoek and Bray (1981) charts provide only an approximate
guide; therefore the comparisons between the Hoek and Bray charts and SLIDE, has
shown that similar results can be derived from each method, although importantly
SLIDE predicts a more exact FOS for the detailed groundwater conditions that were
entered. The comparison has also highlighted a limitation of the Hoek and Bray (1981)
charts, in that they assume failure is through the toe of the slope.

Importantly for the case example presented, groundwater has a significant effect. Both
empirical (Hoek and Bray, 1981 charts) and numerical (SLIDE) approaches,
demonstrate that a critical situation occurs when the phreatic surface is as presented in
circular failure chart 4, within Hoek and Bray (1981). Throughout the analyses outlined
within this section, a particular mass strength has been used, which was derived using
the FracMan-ELFEN approach. The following section analyses the selection of this
particular mass strength, with comparison against a slope mass strength derived
through an empirical-based technique.
221
CHAPTER 8: Case study slope 6 (large slope modelling)

8.5 Limits for slope height and angle from different approaches
It has been demonstrated via slope stability charts, limit equilibrium and finite element
modelling, that the FracMan-ELFEN derived mass strength for the Hoptonwood
Limestone, is too strong for failure through a dry large (1000m) steeply inclined (55°)
slope. However this is dependent on the failure criterion that is used to derive the rock
mass strength and the axial strains at which the strengths are taken.

Two mass strengths were derived using the Hoek-Brown criterion, one set fitted to
peaks in σ1 occurring up to 1.5% axial strain within the ELFEN biaxial models, and the
other at peaks occurring at up to 2.5% axial strain. Only the strength from the 1.5%
axial strain data set, gave a FOS of <1 from a limit equilibrium analysis of a 55° 1000m
slope (see Table 8.12). The mass strength derived using the Mohr Coulomb criterion
appears to provide a better representation than the Hoek-Brown when dealing with
large confinements, as shown in Figure 8.10. Therefore only the Mohr-Coulomb mass
strength was used within the finite element method; both this and the Hoek and Bray
(1981) slope stability charts, consistently show the equivalent continuum to be too
strong for failure to occur through the mass.

What follows is a brief discussion on the various mass strengths derived using the
FracMan-ELFEN approach, and the FOS that some of these produce when using
different analysis methods for particular slope angles. There is a further section
detailing the slope mass strength derived from using an empirical method, the
Generalised Hoek-Brown criterion, and finally comparisons between the two
approaches.

8.5.1 Analysis of downgraded mass strength using FracMan-ELFEN


Approach
A range of Hoek-Brown mass strengths were derived for the downgraded Hoptonwood
Limestone, depending on the axial strain at which the mass strength was taken and the
subsequent failure envelope formed. A lower mass strength was derived when the
envelope for the mass strength was formed by peaks in axial stress that occur at up to
1.5% axial strain, which could be interpreted to represent the onset of failure.
Importantly as presented in Table 8.12, SLIDE showed that a critical point was reached
within the Hoek-Brown mass strength, within this weaker mass (derived from peaks up
to 1.5% axial strain), when the slope angle exceeds 50° within a dry 1000m slope
composed entirely of vertically (ELFEN model RE) aligned Middleton DFN.

222
CHAPTER 8: Case study slope 6 (large slope modelling)

During Section 8.3.2, it was suggested that Mohr-Coulomb criterion is more appropriate
when modelling the considerably higher strength that is revealed at high confinements,
as illustrated in Figure 8.10. Also at this point it was decided that the actual peaks in
vertical stress, (occurring at up to and exceeding 2.5% axial strain), would be used to
derive Mohr-Coulomb mass strength. Consequently a range of mass strengths were
presented to reflect the ELFEN biaxial tests; all of these were considerably stronger
than the Hoek-Brown derived ELFEN mass strengths. This also showed that the Mohr-
Coulomb mass strengths ranged considerably:

 From a low c and a high  (c = 164kPa,  = 58°) derived from the envelope
fitted over the biaxial tests where σ3 = 0, 1, 2 and 4MPa;
 to the higher c and lower  (c = 876kPa,  = 53°) derived from the biaxial tests
where σ3 = 0, 0.5, 1 and 2MPa. Note that as presented in Figure 8.14, it was
found that when the mass strengths with higher  are modelled in SLIDE, a
shallower circular failure surface results (as expected, and discussed 2.2.5).

The mass strength derived from the ELFEN biaxial tests where the σ3 = 0, 1 and 2MPa
provided values that were within the middle of the range suggested by the ELFEN
biaxial modelling. Comparisons are presented within Table 8.16.

These comparisons show that when the mass strength is derived using the Hoek-
Brown criterion, the resulting mass strength is considerably weaker than its equivalent
Mohr-Coulomb mass strength taken from the same range of confinements and axial
strains. In addition to the Hoek-Brown and Mohr-Coulomb comparable mass strengths,
an extra Mohr-Coulomb mass strength is presented that is considered the most
accurate, as it excludes the strength from the ELFEN biaxial test where σ3 = 0.5MPa.

The σ3 = 0.5MPa result is excluded from the envelope, due to the fact that when it is
included, it causes a slight flattening of the envelope (decrease in mass ) and
increase in the intercept (mass c); this alters the slope mass strength slightly, as
presented in Figure J.7 within Appendix J.4. It is recommended there should be further
investigation into the mass strengths that can be derived from using specific
confinements.

Although a Mohr-Coulomb failure envelope is considered linear, in its true form it is


curved, therefore the values of c and  obtained from a Mohr-Coulomb envelope can
change with confining stress. It is the range of biaxial confinements that requires

223
CHAPTER 8: Case study slope 6 (large slope modelling)

further research, providing evidence to the amount of appropriate confinement and the
range of potential strengths that can be derived.

Importantly this extra Mohr-Coulomb (more accurate) mass strength is used in both
SLIDE, Phase2 and the Hoek and Bray (1981) circular failure charts, with the result
from the charts also listed in Table 8.16 (the calculus for the circular failure charts is
presented in Table 8.17). This demonstrates that SLIDE and circular failure charts
agree, with only a 10% difference.

Table 8.16: Comparisons between different criteria used to derive and analyse the mass strength from the
FracMan-ELFEN approach.

Approach used Resulting FOS from SLIDE for a


Factors related to Rock mass
to assess rock 1000 m slope with overall slope
derivation of mass strength
45 50 55
slope mass angle of:
strength parameters
strength
Mass strength taken
from
mb = 3.95
up to 2.5% axial strain 1.3 1.2 1.1
s = 0.0076
in tests where σ3 is
0, 0.5, 1 and 2 MPa
Mass strength taken
from
 = 53
c = 876 kPa
SLIDE up to 2.5% axial strain 2.0 1.8 1.6
in tests where σ3 is
0, 0.5, 1 and 2 MPa
Mass strength taken
from
 = 54
c = 577 kPa
up to 2.5% axial strain 1.9 1.7 1.5
in tests where σ3 is
0, 1 and 2 MPa

 = 54
Hoek and Bray c = 577 kPa
Chart 1 (dry conditions) 1.8 1.6 1.4
(1981) Charts

The implementation of Phase2 permitted the use of the SSR method. This
demonstrated that for a 55° 1000m slope with a mass strength c = 577kPa and
 = 54°, the critical SRF was 1.31; therefore Phase2 calculated that the FOS = 1 when
mass strength c = 440kPa and  = 46°. These mass strength properties were then put
back into SLIDE, to give a FOS of 1.1.

224
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.17: Calculus and consequent FOS from the Hoek and Bray circular failure charts (1981)
for a 1000m slope (where the unit weight of Hoptonwood Limestone = 26kN/m3); note that in this
case F also refers to factor of safety (FOS).

Value from Value from FOS from FOS from

tan  tan 
x-axis y-axis x-axis y-axis

x y F F
Slope Angle

HF Hx
c c
F y
55° 0.016 1.00 1.4 1.4
50° 0.014 0.87 1.6 1.6
45° 0.012 0.77 1.8 1.8

Note that as reported in Section 8.4.2, this SRF was mesh-dependent, with a coarser
mesh predicting the slope to be stronger than with a fine mesh. As a consequence of
this the coarser model actually predicted a higher SRF, which corresponded with the
FOS that SLIDE predicted. This suggests there needs to be further research into
mesh-dependency and the difference between limit state (SLIDE) and finite element
(Phase2) solutions for FOS.

Finally, the influence of groundwater upon the ELFEN-derived mass strength was
analysed with a comparison again between SLIDE and the Hoek and Bray (1981)
circular failure charts, in Section 8.4.3. This showed a difference of up to 20% between
the two methods. This was explained by differences in the location of circular failure
plane, with SLIDE predicting a smaller circle (with a lower FOS when wet) than the
larger critical circular plane (with a higher FOS in wet cases) predicted by the circular
failure charts. This could be due to non-exact matches between the curves for phreatic
surfaces within SLIDE and the illustrations given by Hoek and Bray (1981). Therefore
this aspect is also worthy of further research.

8.5.2 Analysis of mass strength using RocLab-SLIDE Approach


As discussed in Section 2.5.1, one of the quickest (empirical-based) methods to derive
an estimate for rock mass strength, is to use RocLab (Rocscience, 2008). Section
8.3.3 presents the results of this approach for the Hoptonwood Limestone. The
strength of the downgraded Hoptonwood Limestone, derived using RocLab, has been
analysed for different slope angles using SLIDE (Rocscience, 2008); the results for this
study are presented in Table 8.18.

225
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.18: Rock mass strength for the downgraded material, derived directly from intact strength
using the empirical Generalised Hoek-Brown Criterion.

Resulting FOS from SLIDE for a 1000 m slope


UCS
45 50 55
GSI mi with overall slope angle of:
(MPa)
60 12 1.1 1.0 0.9
10
70 12 1.2 1.1 1.0
60 15 12 1.2 1.1 1.0
70 12 1.4 1.3 1.2

GSI for Hoptonwood Limestone


As reported in Table 8.8, Pine et al. (2006) suggested that a typical GSI for the
Hoptonwood Limestone is between 60 and 70; consequently Table 8.18 reports on the
mass strength within this range. However, Elmo (2006) reported values above and
below this range, in relation to the fracture intensity (P21). Elmo (2006) derived a GSI
and a corresponding mi from the biaxial testing upon two Middleton 2.8m by 7m pillars,
one with a P21 of 1.8m/m2 (RB) and the other with a P21 of 2.6m/m2 (RA), finding that:

 The mass strength of a pillar with a P21 of 1.8m/m2 conformed to a GSI range
of 70 to 80;
 Whereas a higher P21 of 2.6m/m2 matched with a GSI range of 40 to 50.

The latter result (from modelling on section RA) is comparable to the data obtained
from the research in this thesis, where RE and RC have been simulated both with a P21
of 2.6m/m2. This comparison is made in Figure 8.21, where suitable RocLab-derived
envelopes are used to suggest an appropriate GSI and mi.

Figure 8.21: Comparison of data from RC and RE biaxial tests with the results that Elmo
(2006) found from testing RA, which has the same fracture intensity.

226
CHAPTER 8: Case study slope 6 (large slope modelling)

Figure 8.21 demonstrates the significant difference in the mass strength derived within
this thesis and that suggested by Elmo (2006) for the same fracture intensity; note that
all models have the same parameters and have been run on the same version of
ELFEN, all which is outlined in Section 8.3.1. This demonstrates how the GSI system
lacks ability to assess the significant influence of fracture orientation, as all of the
models presented in Figure 8.22, can be assigned the same value of GSI. However,
as presented in Figure 8.21, mass strength of these are considerably different; it could
be suggested that RA perhaps has more unfavourably orientated discontinuities, with
less persistent subvertical fractures than in RC and RE, there is the lower likelihood of
a strong interior developing. Instead the shorter subvertical fractures are less
favourable as they can link with the 45° fractures, amounting in a weaker mass
strength.

Another important issue that may contribute to the weaker mass strength noted within
RA, is the fact that the mass strength that Elmo (2006) reports for RA, has been
derived from lower strain rates. Elmo (2008, personal communication) advised that
when modelling the RA section, clear peaks in σ1 occurred at between 0.8 to 1.6%
axial strain; therefore the failure envelope suggested by Elmo (2006) is likely to
represent the onset of failure which is perhaps suitable for pillar strength but not slope
strength, as discussed in Section 8.2.2. This suggests another limitation of the
RocLab-GSI approach, which is the fact that strain dependence cannot be considered.

Figure 8.22: DFN within models RA, RC and RE, which all have the
same fracture intensity (P21) and GSI. However fracture orientation
differs causing a significantly weaker mass strength in RA.

Finally, Figure 8.21 demonstrates that the horizontal model is particularly strong at high
confinement, which could suggest a high GSI and mi. This demonstrates the inability
of the GSI to consider fracture orientation. Also RocLab recommends the limit of 15 for

227
CHAPTER 8: Case study slope 6 (large slope modelling)

the mi of crystalline limestone; however to achieve an envelope for the horizontal


models that is closer to the lower strength at low confinement, the mi would also have
to be larger than 15. This again identifies the inability of the RocLab approach to
consider high strength at high confinement.

Figure 8.23 presents a comparison of the downgraded modelled mass strengths,


against RocLab-generated Generalised Hoek-Brown strength envelopes. This
confirms a high GSI for RE. Figure 8.23 also shows the significantly higher mass
strength of the horizontally aligned Middleton DFN, which cannot be considered when
σci = 12MPa and mi is limited to 15 for limestone, as suggested by RocLab
(Rocscience, 2008).

This again reiterates the fact that the RocLab-GSI approach cannot consider fracture
orientation, but also suggests that the empirical-based approach cannot address the
ratio between the GSI and σci. Purely continuum models can provide an accurate
representation of mass strength when both the GSI and σci are low, however when the
GSI is considerable, as with the Hoptonwood Limestone, discontinuum modelling is
more suitable and consequently the continuum-based RocLab-GSI approach can be
considered inappropriate (as can be interpreted from Figure 2.8).

Figure 8.23: Data from downgraded FracMan-ELFEN tests. The horizontally aligned
Middleton DFN, presents a high mass strength that cannot be considered using mi
suggested in RocLab.

The equivalent continuum approach presented within this chapter, has demonstrated
that unlike the RocLab-GSI approach, a high GSI:low σci situation can be represented.
Consequently the significantly higher shear strength of a dominantly vertically fractured
228
CHAPTER 8: Case study slope 6 (large slope modelling)

mass, which could exist in the base of a large slope failure, appears to be unaddressed
within the RocLab-GSI empirical based method.

Despite the limitations of the RocLab-GSI approach that are discussed, an improved
match between the FracMan-ELFEN derived mass strength and the RocLab mass
strength, can be achieved when a higher GSI is adopted. Both Figures 8.22 and 8.23
have demonstrated that the GSI for the Middleton DFN could exceed the 60-70
previously suggested. Consequently the results listed in Table 8.18 can be re-
interpreted with a higher GSI, using SLIDE, to give a slightly higher empirically-derived
slope mass strength, as presented in Table 8.19.

Table 8.19: Rock mass strength for the downgraded material, derived directly from intact strength
using the empirical Generalised Hoek-Brown Criterion with a higher GSI.

Resulting FOS from SLIDE for a 1000 m slope


UCS
45 50 55
GSI mi with overall slope angle of:
(MPa)
80 10 12 1.4 1.3 1.2
80 15 12 1.6 1.5 1.3

8.6 Summary and Conclusions

8.6.1 Selection of DFN


The material selected, the Hoptonwood Limestone, has revealed a mass strength that
is too strong to permit circular failure in a dry slope up to 1000m deep at a reasonable
slope angle (55°). However, as discussed in Section 2.2.5, large scale rock mass
failure within an open pit can be complex and potentially largely along discontinuities
rather than mass controlled. This makes the derivation of an equivalent continuum, for
consideration of circular failure, problematic and possibly of limited value.

The Middleton DFN, created by Elmo (2006), was chosen because of previous ELFEN
uniaxial and biaxial modelling. This also provided a model with considerable fracture
intensity and an intact rock strength that is equivalent to the common reasonably
competent lithologies that appears to be prevalent within the present-day large-scale
open pits. It is unlikely that the intact strength is too high, or the DFN is not fractured
enough, to allow the failure of a large slope. Instead it is the large-scale failure-type
that needs more consideration.

229
CHAPTER 8: Case study slope 6 (large slope modelling)

8.6.2 ELFEN Biaxial Modelling


Biaxial modelling within ELFEN is a lengthy process. The derivation of mass strength
has shown itself to be open to many variables. Primarily, as discussed by Starfield and
Cundall (1990), discontinuum analysis can be very sensitive to change. As ELFEN is a
comprehensive hybrid approach, there are many aspects which can alter the result of
the simulation. The primary difficulties experienced during the ELFEN biaxial modelling
are listed below:

 Different results with different analyses versions, which would have required
detailed calibration to acquire an accurate representation; consequently a non-
current version of ELFEN (analysis 3.3.31), had to be used for analysis.
 Using this version of ELFEN, no simulations could be restarted, as this would
otherwise impart dynamics and cause the model to prematurely fail.
Subsequently in one case, analysis could not be ceased for approximately
14 days.
 The detection of the actual peak and yielding behaviour of models was difficult,
especially with high confinements, which prevented clear fracture planes
developing, but signified the importance of strain magnitude relative to the
failure environment.

The importance of strain magnitude was reiterated when comparing the relevant biaxial
testing that Elmo (2006) presents, against the results from highly strained models from
this thesis. This was shown in Figure 8.21, where the mass strength suggested by
Elmo (2006) is considerably weaker than the mass strength derived within this thesis.
This is due to the fact that Elmo (2008, personal communication) derived the mass
strength of section RA, from clear peaks in σ1 that occurred at between 0.8 to 1.6%
axial strain; therefore the failure envelope suggested by Elmo (2006) is likely to
represent the onset of failure which is perhaps suitable for pillar strength but not slope
strength.

8.6.3 Modelling of the Equivalent Continuum


Three different methods have been used to analyse the strength of the equivalent
continuum derived from FracMan-ELFEN testing. Two of these techniques (circular
failure charts and SLIDE) reveal a FOS for a particular slope design; consequently
comparisons can be drawn between the two methods. The results and benefits of the
three approaches are summarised in the following table.

230
CHAPTER 8: Case study slope 6 (large slope modelling)

Table 8.20: Summary of methods, advantages, disadvantages and comparisons between each approach
used in the analysis of the large slope case study.

strength derivation

Method of analysis
Representation of
Method of mass

mass strength
Comparison
Further
between numerical
Advantages Disadvantages recommend-
and empirical based
-ed work
mass strength

Can only accept Investigate if


Mohr-Coulomb there is a
Hoek and Bray failure
(FracMan-ELFEN)

mass strength particular


Quick and Generally good
Mohr-Coulomb

(not Hoek-Brown) mass


Numerical

approximate comparison to
cohesive
charts

‘staging post’ Limited sensitivity empirically-derived


to cohesive strength,
for further, slope mass strength
strength below which
more detailed (when assessed in
the circular
analysis SLIDE)
Failure must be failure charts
through toe of are
slope inaccurate

Mass strength
Coulomb

Investigate
Mohr-
(FracMan-ELFEN)

dependent on the Strong slope mass


- influence of
biaxial strength
Numerical

selection of
confinement
biaxial
Highly variable confinements
Can directly
due to inability to Lower slope mass on Hoek-
SLIDE
Brown
Hoek-

take Hoek-
represent high strength than Mohr- Brown mass
Brown
strength at high Coulomb strength
properties (no
confinement
need to derive
Mohr-Coulomb
Empirical
(RocLab)

Brown
Hoek-

properties High dependency


Low slope strength -
using σ3max on the GSI
method)
Detection of
critical slope
mass strength
via SSR
No failure within Application of
method
downgraded FracMan-
Comparison of
Hoptonwood ELFEN
σ1 σ3 pairs is A direct FOS for
(FracMan-ELFEN)

limestone (too approach to a


Mohr-Coulomb

possible (at specific scenarios strong). Thus σ1 σ3 weaker rock


Numerical

locations that cannot be pairs could not be


Phase2

failed within the mass,


derived compared
equivalent allowing
(although SRF
continuum) with further
can derive detect
those predicted comparisons
strength for
in biaxial tests between
FOS = 1) 2
Phase and
Stress-path modelling initial biaxial
Stress-path
only performed on modelling*
monitoring is
equivalent mass that
possible during
had been
modelling
downgraded by SSR
stages
method

* In particular it is the angle between the resulting σ1 and circular failure plane, at
different confinements, which would be of interest if failure was detected in the
231
CHAPTER 8: Case study slope 6 (large slope modelling)

equivalent continuum. This was demonstrated via the use of the SSR process within
Phase2, and is presented within Figure 8.19b; although in this case, no comparison can
be made to the initial ELFEN biaxial modelling. This is because the failure predicted in
Phase2 was through a weaker equivalent mass than was initially used within the
ELFEN biaxial simulations.

8.6.4 Final Summary


The comparison between the empirically and numerically derived slope mass strength
has highlighted many issues that cannot be considered within the empirical-based
approach used:

1. There is no consideration of the fracture orientation within the GSI assessment;


consequently the empirical-based approach cannot assess the significantly
stronger region within the toe of a slope, due to shear against a dominantly
vertically fractured mass during circular failure.
2. The RocLab-GSI equivalent continuum approach is limited when GSI is high,
also mass properties are ultimately based on the Hoek-Brown criteria. This
criterion cannot represent the high strength at high confinement in this case;
instead a discontinuum based assessment may be more appropriate, such as
the FracMan-ELFEN approach.
3. The FracMan-ELFEN approach has illustrated that significantly different mass
strength can be derived from low and high strains. Particular strain rates
cannot be addressed using the empirical-based approach, which may be
important in different failure environments.

In addition, careful consideration is needed during the derivation of representative


mass strength parameters from the FracMan-ELFEN biaxial tests. The mass strength
properties are not only a function of the strain magnitude from which they were derived,
but also there is a significant dependency on the confinement within the biaxial tests,
which were used to create the consequent failure envelope. It was found that in the
case of the downgraded vertically fractured (RE) section, it was most appropriate to
use results from σ3 = 0, 1 and 2MPa; further research into this is necessary with
consideration of slope scale. The same conduct was carried forward for the
horizontally fractured mass. This aspect requires further investigation, with reference
to the σ3max that is likely within the slope being considered. It would be of great value, if
the FracMan-ELFEN approach was used as a back-analysis tool, to determine the
accuracy of the method in both a mass and a mixed mass-discontinuity rock slope
failure.

In terms of efficiency of the ELFEN simulations during the FracMan-ELFEN approach,


it has been shown that simulations are sensitive to parameter variation, limited to a
232
CHAPTER 8: Case study slope 6 (large slope modelling)

single processor and are consequently slow. Within biaxial tests the time-step is
reduced significantly by a low mesh quality, due to the problematic meshing routine
generating small ‘sliver’ elements within acute discontinuity/geometry intersections.

Importantly the empirically-derived mass strength is lower than the numerical-derived


mass strength. Also the FracMan-ELFEN approach has provided a highly variable
mass strength, which could be considered more realistic when analysing the complex
behaviour of a fractured mass. It is therefore recommended that a Monte Carlo
method should be considered in future biaxial modelling.

In the case of Hoptonwood Limestone, it appears that an equivalent continuum


approach alone may be too simple. Instead mixed structural and mass-controlled
slope failure needs to be considered, which may involve more scales of biaxial
modelling to fully characterise a rock mass. On the smallest scale, the intact strength
of the rock-bridges needs to be accurate; with the Hoptonwood Limestone this was too
strong for large-scale circular failure within the equivalent continuum. As a result, the
intact strength of the Hoptonwood Limestone was degraded to create a new rock mass,
which retained the DFN originally created for Middleton Limestone Mine. Biaxial
testing was then performed on this downgraded rock mass, arriving at an equivalent
continuum within which large-scale slope failure was modelled, using SLIDE and
Phase2.

It is recommended that the intact strength of the Hoptonwood Limestone in the biaxial
sections is reviewed. This may involve a smaller scale of biaxial modelling, to give
more detail on the rock-bridge strength between discontinuities. Consequently this
would influence the equivalent continuum strength within the medium-scale biaxial
models that were simulated within this Chapter. In the case of an actual slope, large-
scale structures should have also to be considered, which may involve another larger-
scale of biaxial modelling, to further degrade the equivalent continuum. This would
require assessment of the representative elementary volume (REV).

Large slope-scale discontinuum modelling is obviously preferable over an equivalent


continuum approach, where it is otherwise necessary to downgrade rock mass strength
via biaxial modelling on several scales. Large-scale discontinuum modelling would
allow discrete representation of large structures in combination with an appropriate
equivalent continuum strength for the mass. However, for large-scale discontinuum
modelling to be viable, there needs to be a significant advance in the computational
speed of the present day comprehensive codes that are available.

233
CHAPTER 9: General Discussion

CHAPTER 9: General Discussion

9.1 Introduction
This Chapter reviews the content of this dissertation and specific findings from the
numerical modelling methods, which have been used.

An investigation of the ability of different numerical methods to analyse slope stability in


fractured rock slopes. In particular one comprehensive approach, ELFEN, is examined
with comparisons against various less complex limit equilibrium methods and a finite
element method (Phase2). During the time span in which the research was conducted
a new version of ELFEN, which included a groundwater module, was developed. This
was tested through three of the six case study slopes presented, proving that it is
effective in reducing discontinuity-based normal stress, relative to pore pressure. With
the ability of ELFEN to insert mode I fractures, progressive failure was simulated as a
result of a rise in groundwater.

9.1.1 Overview of thesis content


Chapter 1 gave an introduction, describing the context of the research within the
discipline of geomechanics, outlining subsequent research aims and objectives.
Previous work with ELFEN was briefly reviewed, with further more in-depth discussion
on the application of ELFEN to rock mass strength, during Section 2.5.2.

Six case studies were presented as an approach to the research aims outlined in
Chapter 1. Chapter 2 was a literature review recording detail of respective areas that
had to be considered during this thesis. Chapter 3 introduced particular aspects that
have to be considered when developing a model of a fractured rock slope within a
dynamic fracture-based code. Consequently a staged modelling approach was
presented from which one can accurately simulate failure of a fractured rock slope, as
a result of a particular trigger process, using a dynamic fracture-based code.

In Chapter 4 difficulties were experienced when comparing the behaviour of a simple


planar failure model in ELFEN with specific limit states from analytical solutions, as
reported in Chapter 4. This highlighted the sensitivity of ELFEN, to numerical
parameters, with calibration of certain properties required to obtain apparently correct
behaviour. Finally this was achieved within an ELFEN model, comparing well with limit
state solutions; this was within a model that used the staged modelling regime

234
CHAPTER 9: General Discussion

suggested in Section 3.1. Later within Chapter 4, a tension crack was introduced into
the plane failure model to provide further examination of the solution in ELFEN when
groundwater is implemented. Both this tension crack and the non-tension crack model
formed case study slope 1; one of the six different case study slopes of increasing
scale that are presented throughout Chapters 4 to 8.

Case study slope 1 showed that the comprehensive numerical approach was
particularly sensitive to mesh density, and the consequent shape of the phreatic
surface. At first a relatively coarse mesh was used providing a simplified phreatic
surface in a relatively rapid simulation. However, in situations where the solution was
highly dependent on the pore pressure acting upon the discontinuity, unity with the limit
equilibrium method could not be achieved with the simplified phreatic surface, most
probably due to excess pore pressure within the discontinuity. Therefore the phreatic
surface were modified in ELFEN, to attempt to achieve similar uplift forces due to pore
pressure, between ELFEN and the limit equilibrium method. This improved the
solution; however behaviour with ELFEN then appeared to over-estimate stability, most
probably due to insufficient pore pressure within the discontinuity. Consequently it was
suggested that a finer mesh density, and associated longer runtime, is required to
improve the accuracy of the phreatic surface.

In the same way that the optimised mesh-density within ELFEN is important for the
simulation of pore pressure, it is also highly influential in the development of shear
strain and consequent mass-controlled failure. This was evident in both ELFEN and
Phase2 simulations of case study slope 2, the Hutchinson Joss Bay failure presented in
Chapter 5. Comparisons of different modelling approaches showed that the back-
analysed mass strength was similar, to that from the slice based limit equilibrium, to the
more accurate Phase2 and ELFEN models. In addition it was found that a pre-existing
tension crack within a Phase2 model significantly influenced the back-analysed rock
mass strength. It is recommended that to confirm this there should be further review of
the joint stiffness applied to the tension crack within the Phase2 model, and
construction of simplified ELFEN and Phase2 models for comparison against RocPlane
and Plane_failure. In addition the Microsoft© Excel spreadsheet Jacob (Pine, 2006a),
should also be considered in future analysis.

In order to analyse the influence of groundwater within a similar model of a chalk cliff, a
more discontinuity controlled failure had to be developed. In this case a step-path
model was created (case study slope 3), providing the required criteria to test the
current feature of the newly developed groundwater module within ELFEN. During the

235
CHAPTER 9: General Discussion

development of this model, dynamic-based problems highlighted the sensitivity of the


ELFEN code, and the importance of modelling in stages. In this case modification of
failure geometry ensured the desired failure response, without detailed investigation of
rock-bridge strength and mesh density, which would otherwise be necessary.

Aspects of dynamic activity within ELFEN, also proved problematic during the
simulation of the larger case study slope 4, based upon the failure at the Delabole slate
quarry reported in Chapter 6. The location of rock-bridges and extent of fracture
network were investigated, with progressive development of the model until an
apparently realistic failure mechanism and resulting block interaction were represented.
Groundwater was introduced into primarily a simple planar failure model, and finally a
more complex model with a rock-bridge. To achieve failure as a result of a rise in the
phreatic surface, dynamic-based failure had to be addressed, as discussed in
Section 9.2.

Case study slope 5 was introduced in Chapter 7, where a variety of limit equilibrium
methods and also two Phase2 models were used to investigate rock mass strength
within the toe of an intermediate-scale failure. This showed that most models returned
similar results, with one exception, which had to be modified to achieve realistic
behaviour. Most of the different limit equilibrium methods were restricted in their ability
to consider certain situations during the sensitivity analyses; this is due to the
simplifying assumptions implicit in such models. Finally, recommendations were given,
outlining the further detail that is required to develop a more advanced numerical
model, which may potentially give a better representation of rock mass strength within
the toe region of the failure.

Rock mass strength within the final large-scale slope case study was extensively
reviewed within Chapter 8. The numerical-based method (FracMan-ELFEN approach)
was used to generate equivalent continuum strength properties representing a
fractured mass. This fractured mass and the subsequent initial strength model was
based on the work of Elmo (2006), who completed mapping, statistical fracture
generation and pillar modelling within the Hoptonwood Limestone exposed within
Middleton Mine in Derbyshire, UK. It was discovered that FracMan-ELFEN approach
showed high strength at high confinement, which meant that a linear failure criteria
(Mohr-Coulomb) provided a better representation of the strength data, than a non-
linear method (Hoek-Brown), in this particular case.

236
CHAPTER 9: General Discussion

The first equivalent continuum that was derived for the Hoptonwood Limestone was too
strong to allow the consideration of mass-controlled circular failure. Therefore intact
strength was quartered and the biaxial sections were re-modelled, returning a lower
mass strength that was used to show how the FracMan-ELFEN process could be
completed in a large slope. Less complex limit equilibrium and finite element methods,
were then used to study the same large-scale slope behaviour. In addition a zoned
slope was considered where the axes of principal stress, within biaxial sections, were
rotated to allow simulation of an appropriate shear strength considering the dominant
discontinuity orientation in relation to the path of the circular failure. This gave
equivalent continuum properties that exceeded the strength which is applicable to
mass controlled failure. Thus an ‘un-zoned’ slope was mainly used during the analysis
of stress and subsequent comparisons. Importantly this emphasised that there is no
consideration of the fracture orientation within the Geological Strength Index (GSI)
assessment, and therefore the empirical approach is (as originally designed)
specifically limited to a situation where there is no dominant discontinuity orientation
that may influence the path of a circular failure.

Comparisons were made against the numerical-based rock mass strength and
empirical-based methods, showing the latter to return lower equivalent continuum
strength properties than the former. An interpretation could be that the continuum
approach is more accurate (and less conservative) than an empirical-based approach.
Further application of the FracMan-ELFEN approach to a real large slope failure is
recommended, as outlined in Section 10.3.

The following Section reviews specific advantages and disadvantages that have been
found with ELFEN. Subsequently Section 9.3 makes comparisons against ELFEN and
other numerical approaches.

237
CHAPTER 9: General Discussion

9.2 ELFEN capabilities and limitations


Throughout this thesis it has been shown that ELFEN has unique features and a user
interface that, although complex, shows progression towards a more user-friendly
software than the programming that is required within alternative numerical codes to
achieve similar results. However, there are many features and modifications that are
applicable to fractured rock slope models, which are not detailed in the current ELFEN
manual. This thesis has presented all of the approaches used; also a staged approach
has been developed, by which a numerical model of a fractured rock slope should be
developed. Hopefully this provides guidance for future numerical modellers when
using a comprehensive dynamic fracture-based code for slope simulations.

This staged approach, and the restart feature within ELFEN allows the interrogation of
both rock mass and discontinuity strength within a slope. In order to provide an
accurate representation of strength, the many numerical parameters that are required
(all summarised in Appendix A), need calibration as each of the parameters can
independently have a significant effect on the solution. Consequently, throughout
Chapters 4 to 6, trial models have been discussed where some of the parameters and
subsequent behaviour within ELFEN is calibrated against known solutions. As the
step-path simulations demonstrated, during the modelling of case study slopes 3 and 4,
the use of alternative methods is not always possible; instead one has to rely on a
dynamic-based approach and try to calibrate rock-bridge strength with the desired
behaviour.

Ideally, before simulating a fractured rock slope, a detailed set of intact strength and
fracture network data would have been collated. This would allow calibration during
detailed back analysis of the model, using the limit point suggested from pre-failure
studies. This is not always possible, instead in the case of this research, the more
complex step-path models (within Chapters 5 and 6) were developed from preceding
more simple models that were calibrated with analytical and finite element method
(FEM) numerical approaches. Therefore it is always desirable to constrain the
conditions within a simple ELFEN model first, with comparisons against less complex
software; once this has occurred one can then develop the model further, introducing
rock-bridges and extending the fracture network if necessary.

Throughout this research, it has been found that ELFEN is sensitive to dynamic loading
and energy dissipation. To ensure the stability of a slope model prior to the appropriate
trigger process, kinetic energy has to be minimised; as a response to this the staged

238
CHAPTER 9: General Discussion

process was developed (as presented in Section 3.1). In particular, difficulties with
dynamic-related failures were experienced in situations where progressive failure
through rock-bridges is required. When the trigger of a progressive slope failure is a
rise in groundwater, a model which is stable in the dry and not in the wet has to be
achieved. Particular examples were given during the modelling of:

 An example chalk cliff section, within Section 5.3


 An actual failure at Delabole slate quarry, Cornwall, outlined in Sections 6.3.2.

The modifications made during the Delabole modelling, summarised in Table 6.5,
perhaps provide a more proactive way of arriving at a suitable simulation, than in the
theoretical Chalk modelling. Instead during the Chalk step-path modelling, modification
of geometry occurred until a suitable situation was found.

9.3 Brief comparisons between ELFEN and other


comprehensive numerical approaches
This section makes comparisons between ELFEN and Phase2 from the experience
gained throughout this research. In addition further comparisons are made, with an
alternative discontinuum method which has fracturing capabilities, Particle Flow Code
(PFC). Although this was not used during this research, advantages have been
identified within literature.

9.3.1 Phase2
Phase2 provides a more simplified approach than in ELFEN. However, there is
considerably less control on loading schemes and post-processor monitoring. The
ability of Phase2 to simulate complex interactions from discontinuities within a model is
restricted, due to the continuum FEM upon which Phase2 is based.

Despite this, it has been demonstrated that, as with ELFEN, Phase2 can be applied to
both discontinuity and mass controlled slope instabilities. During this thesis, only
2
Phase models with limited discontinuity control have been presented. The Delabole
failure (case study 4) was simulated within Phase2, although this model was not
successful enough to present; therefore it is suggested that there could be further
research into the ability of Phase2 in situations where there is a more extensive fracture
network and ‘blocky’ behaviour.

239
CHAPTER 9: General Discussion

Chapter 5 reported on the use of Phase2 in the analysis of the Hutchinson Joss Bay
failure (case study 2). There is an automated Shear Strength Reduction (SSR) method
within Phase2 (see Section 2.3.4). This provided a rapid way of back-analysing the
mass strength of the chalk. Also the SSR process within the Hutchinson Joss Bay
Phase2 model (without the pre-existing tension crack), demonstrated that Phase2 can
potentially predict the location of tension cracks, as a result of progressive shear. This
was also the case within ELFEN, although there is no automated SSR function
available. Instead ELFEN provided the opportunity to examine the influence of model
unloading, by inspection of stress conditions at particular points, at a level of detail
which Phase2 cannot provide.

In addition to mass controlled instability, Phase2 can be used to evaluate resistance


provided by discontinuities. Chapter 7 presents a detailed investigation of the mass
strength of the rock-bridge at the toe of a potential active-passive-type failure. Two
models were created:

1. A model with a discontinuity representing the rock-bridge, upon which the


degree of yield was noted.
2. A SSR model where strain was permitted through the localised mesh which
represented the rock-bridge at the toe of the failure.

Both of these models returned similar mass strengths for the rock-bridge. However, it
became clear that the automated SSR method was limited to the second model listed
above. In the first model a manual method was necessary, noting the degree of
yielding upon each of the discontinuities controlling the failure.

Although automated, the SSR simulations of the second model demonstrated how a
full sensitivity analysis cannot be conducted, without manually applying the method, as
discussed in Section 7.3.2. This is because the Shear Strength Reduction Factor
(SRF) cannot be applied to each of the strength parameters (c and ) independently,
preventing independent analyses.

In addition, during the simulations presented in Section 8.4.2, a SSR process was only
conducted with a Mohr-Coulomb strength because it is difficult, when using Hoek-
Brown strength, to calculate the mass properties from an appropriate SRF. This is due
to presentation of the Hoek-Brown Strength criterion as a function of σ3 rather than
shear stress. As outlined in Section 2.3.4, Rocscience (2004) state that it is impossible
to generate close-formed relations, as those that can be created when using linear

240
CHAPTER 9: General Discussion

criteria. Therefore it could be suggested that back-analysis using the SSR approach is
limited to rock masses that can be represented by Mohr-Coulomb and other linear
failure criteria.

Finally, although it is considerably more efficient to use the SSR process, than
manually trying to back-analyse rock mass strength, simulations should be carefully
setup. With fine iterations, even SSR models with specified search areas can have
long run times; for instance it took in excess of an hour to complete 20 iterations within
the zoned 1000m slope presented in Section 8.4.2. During this time the progress of
the simulation cannot be viewed as it can in ELFEN, although there is a useful ‘pause’
and restart feature.

The following section reviews the unique features of ELFEN in comparison to those of
a discontinuous method, PFC.

9.3.2 PFC
As discussed in Section 2.5.2, Itasca and Rockfield are currently the only two
companies identified that have the relevant codes (UDEC/PFC and ELFEN
respectively), commercially available and developed enough to allow accurate
geomechanical modelling of a discrete fracture network (DFN). The primary aim of this
research was to assess the ability of ELFEN to deal with rock slope failures through
blocky rock masses. Therefore PFC or UDEC have not been used due to logistical
constraints and access to up-to-date codes; consequently no direct comparisons can
be made, although from review of literature it is clear that there are certain advantages
and disadvantages associated with each approach.

Throughout this thesis rock mass strength has been represented, by the use of an
appropriate strength criterion. ELFEN models have used either a Mohr-Coulomb,
Rankine Rotating crack or a coupled approach, where Rankine forms a tensile cut off
to Mohr-Coulomb (see Appendix B). Read (2007) describes the fracture criterion basis
within ELFEN as a macro mechanics based approach, whereas in PFC a micro-
mechanics approach is used. This allows PFC models to be detached from an
empirical-basis, without foundation on any particular criterion. This can be viewed as a
particular advantage of PFC over ELFEN; although, it can lead to difficulties in
generating the correct micro-bond strengths in PFC for comparison with macro-
strength effects.

241
CHAPTER 9: General Discussion

As well as the independence of PFC from failure criterion, Read (2007) suggests that
calibration of the uniaxial compressive strength (UCS), Modulus and/or Poisson's Ratio
can occur in PFC model during an initial elastic continuum stage. This is an important
quality that should perhaps be considered within ELFEN, because it allows direct
interrogation and any necessary adjustments of numerical parameters during the
primary modelling stage.

Section 2.3.7 discusses particular features of PFC; like ELFEN, PFC can now
accurately simulate fracture extension through a ‘smooth contact model.’ However it
remains that there is no post-failure representation of any discrete blocks that are
created, when using PFC. The ability to form discrete failure blocks is a feature that is
unique to ELFEN, this allows the modelling of the complete failure process and is
therefore a particular advantage of ELFEN over PFC.

The fracture process within PFC is simulated by the breaking of individual bonds
between particles and coalescence of these to form microcracks. In ELFEN, fractures
can extend through individual elements (intra-element fracturing). As discussed in
Section 2.3.8, intra-element fracturing could be a significant benefit of simulation using
ELFEN; conversely during this research it has been discovered that intra-element
fracturing can cause numerical instability and consequent excessive runtimes.
Therefore this beneficial fracturing mode within ELFEN is possibly limited at present.
Further research is required to verify this, although at present it seems that PFC and
ELFEN could be considered as fundamentally equal in their individual capabilities to
simulate fracturing.

242
CHAPTER 10: Conclusions and recommendations

CHAPTER 10: Conclusions and recommendations for


further work

10.1 Introduction
This Chapter provides final conclusions, with a summary of the contributions to
research. Also recommendations are made for further research within the specific
case studies outlined throughout this thesis.

10.2 Conclusions
The research within this thesis has contributed to the field of numerical modelling in
geomechanics principally by evaluating a comprehensive fracture-based code, ELFEN
(Rockfield, 2008), for the simulation of fractured rock slopes. A variety of different
scales and failure mechanisms were modelled and some case study slopes included
groundwater. This alone has shown that:

 ELFEN is particularly sensitive to dynamics induced by essential modelling


stages, which can cause ‘kinetic instability’ and artificial failure. To avoid this a
staged modelling approach was developed, and is presented to allow the
accurate simulation of failure in fractured rock slope due to a specific trigger
process.
 Calibration of numerical parameters within ELFEN is necessary if one is to
achieve a realistic response. This calibration can occur through comparisons
with intact strength, based on laboratory testing or mass strengths derived
empirically. Alternatively in slope situations, failure mechanisms should be
simplified to allow calibration against less complex limit state solutions.
Subsequently model development can occur in a controlled situation to achieve
an accurate failure mechanism, and represent post-failure deformation.
 The re-start function within ELFEN allows a relatively rapid assessment of the
sensitivity of the solution to different numerical parameters. If calibrated and
kinetically stable, back analysis of the rock mass strength and detailed
investigation of a variety of conditions can be achieved.
 The user interface and documentation supporting the use of ELFEN require
further development. At present, tuition, intuition and experience are necessary
to enable the use of the complex modelling software in a meaningful way.

243
CHAPTER 10: Conclusions and recommendations

Comparisons against other numerical approaches have confirmed that the less
complex, limit equilibrium and finite element methods, can be applied to the
assessment of rock mass strength in fractured rock slopes. A key issue, which the
comparisons demonstrated, is the dependency on mesh density. This is not solely an
issue for ELFEN but all mesh-based approaches. During this research it has become
evident that the optimum mesh density is imperative for the accurate simulation of both
shear strain and the effect of a phreatic surface in ELFEN.

Though the development of three suitable slope models, it has been shown that the
current version of ELFEN can reduce normal stress upon discontinuities relative to
pore pressure. This can directly cause instability, or in two of the case study slopes,
can instigate the progressive failure of rock-bridges and resulting failure of a fractured
rock slope.

Slope mass strength at varying scales has been investigated throughout this research.
Mass strength was observed within a bench-scale failure; rock-bridge strength has
been reviewed within a dominantly discontinuity-based failure at intermediate scale,
and large scale rock mass strength has been assessed by the multi-location biaxial
testing of a fractured mass. In particular the latter investigation of large-scale
behaviour has furthered the development of a numerical-based method of rock mass
strength assessment, thus allowing the derivation of an equivalent continuum from a
fractured mass. This was then considered using a variety of techniques at large-slope
scale. In addition, comparisons were made against a conceivably conservative, and
significantly more limited empirical-based method of mass strength assessment.

10.3 Recommendations
The use of alternative numerical approaches has exposed aspects, which could be
recommended as being beneficial to ELFEN:

 Implementation of an automated shear strength reduction process with


independent application of the reduction factor to allow full sensitivity analysis.
 Improved means of viewing results, using a yield approach as in Phase2
(Rocscience, 2008); this would enable the rapid identification of compressive
and tensile mass and discontinuity related failure.
 Improvement of the restrictive intra-element fracturing scheme within ELFEN,
possibly reducing the dependency on mesh density.

244
CHAPTER 10: Conclusions and recommendations

 Development of ELFEN so that fracture extension and insertion, due to mode II


failure, can be simulated.

The FracMan-ELFEN approach was successful in showing that an equivalent


continuum can be derived for a fractured rock mass. However the high strength
derived from initial testing on the Hoptonwood Limestone, suggested that a mass-
controlled failure is unlikely to occur. By downgrading the intact strength properties for
the Hoptonwood Limestone, a new rock mass was created that still explicitly included
the Middleton DFN, which was more applicable to an equivalent continuum approach.
It is recommended that the next step should be to seek a mass-controlled large-slope
failure case study (similar to the downgraded rock mass used in this thesis), to which
the FracMan-ELFEN approach can be applied.

More competent rock masses with high mass strength require a more discrete,
discontinuous analysis as opposed to an equivalent continuum (FracMan-ELFEN)
approach. The simulation of discontinuity-controlled large-scale failure within a slope
with a detailed fracture network, demands further processing power, as does the
simulation of a fractured rock slope in three-dimensions. Currently a limit has been
reached on the speed of a single processor; therefore it is recommended that the
ELFEN requires parallel processing for such tools. In addition neural networking
should be considered to enable the splitting of component parts of the software,
providing computer power to couple further mechanical processes, extending the
potential of ELFEN.

As previously mentioned, to further test the FracMan-ELFEN approach it is necessary


to back-analyse a large slope failure, ideally within a rock mass where mass controlled
failure occurred. However, as discussed in Section 1.1, the access to large-scale rock
slope failures is restricted, with a lack of case studies. Also the modern-day
development of large open pits is now exceeding the knowledge-base upon which
current empirical schemes are founded. It is recommended that empirical approaches
are updated, with a review of their previous foundation, and a presentation of new
larger-scale slope analyses, onto an expanded scheme to provide an up-to-date chart
of potential case study slope failures. Ideally the development of a web based slope
database would make the recording of large slope failures easier and more accessible.

However, the data from such analyses is rarely released from rival mining companies.
Therefore the mining community needs to encourage the release of data from slope
analyses; it is in their interest as the database would provide a means by which

245
CHAPTER 10: Conclusions and recommendations

improved and more accurate approaches can be developed, increasing the safety and
efficiency of modern-day open pit mining.

Sullivan (2007) suggested that mine slope design is now often ‘performance based’
with emphasis on risk and probability of failure, rather than the use of conventional
factors of safety based on well-developed knowledge. It is possible that a risk-based
approach could be integrated with the FracMan-ELFEN method of slope mass strength
determination. One particular way in which the solution of the biaxial testing within
ELFEN could be improved, is to use Monte Carlo simulations of biaxial tests, deriving a
number of equivalent continuum strength properties, from which a range of strength
can be taken. This would improve the reliability of data and add a statistical
component into the numerical-based mass strength assessment.

Comparisons against empirical based approaches exemplified the fact that the Hoek-
Brown criterion is only for equivalent continuums. The existing guidelines that are
available dictate the applicability of the Hoek-Brown criterion, as indicated in
Figure 2.8. The Middleton Mine DFN is perhaps situated upon the boundary of where
the Hoek-Brown criterion is applicable, having many joints, in which case the criterion
should be used with caution. An initial numerical assessment showed that the rock
mass revealed a stronger tendency and therefore perhaps a more structural-based
assessment is necessary. It is recommended that there should be further research into
the ability of the FracMan-ELFEN approach and the alternative SRM methods, in
competent rock masses, where the fracture network is such that the Hoek-Brown
criterion should be used with caution.

246
APPENDIX A: Important parameters within dynamic fracture-based code

APPENDIX A: Important parameters required within


dynamic fracture-based code

This Appendix outlines important parameters that had to be investigated as part of the
familiarisation with sophisticated methods of numerical modelling. In particular the
requirements within the specific code that was used as part of this thesis, ELFEN, are
used to discuss key areas that have to be considered within a dynamic fracture-based
code.

Firstly analysis within ELFEN can be selected to be implicit, explicit, implicit/explicit or


discrete. Discrete analyses within is used where a combination of discrete and finite
elements is desired, as is the case with the simulations concerned with this thesis;
although it is suggested within the ELFEN documentation (Rockfield, 2008) that multi-
phase projects such as geotechnical simulations with groundwater, should use the
implicit analysis. All analyses with groundwater within this thesis have been conducted
using discrete analysis, on account of the required fracturing ability. Further research
is suggested regarding this aspect.

The implicit function within ELFEN is a technique more appropriate to problems where
fracturing is not required and materials behave in a linear fashion with the solution
seeking to achieve equilibrium at each timestep (Rockfield, 2006). At present all
simulations have been discrete analyses following the excavation of an underlying coal
seam worked example in the manual supplied by Rockfield (2008), and the work of
Elmo (2006).

The explicit solver within ELFEN uses a ‘central difference method,’ which effectively
involves the solver tracking the progress of a stress wave through a domain/element
(Rockfield, 2006). A disadvantage of this method is that it becomes unstable if the
stress wave generated within the model passes over the boundary of more than one
element during each timestep. Consequently a maximum (‘critical’) timestep is
included within ELFEN which prevents this from occurring via consideration of the
wave speed through the material (see Section A.3).

Finally much of the functionality within ELFEN is accessed through the text-based
neutral file (‘neu’ file). This is created upon meshing and must be edited to obtain
some of the features discussed within this Appendix and Appendix B. Note that in
versions of ELFEN that post-date the analysis within this PhD, a new pre-processor

247
APPENDIX A: Important parameters within dynamic fracture-based code

has been developed, which now creates a ‘dat’ file, with the same functionality as the
neutral file, however a slightly different layout.

A.1 Material Properties – Non-linear Criterion


Non-linear Criteria allow elastic-plastic mediums, such as rock, to be modelled with
consideration of the permanent strain (damage) processes that are outlined in
Table 2.1. Lightfoot and Maccelari (1999) note the difficulty, of implementing
constitutive laws for plasticity, within numerical models, especially implicit solution
schemes. At the time when Lightfoot and Maccelari (1999) wrote their report, five
plasticity models had been built into numerical modelling programs: Mohr Coulomb
model, Hoek and Brown model, ubiquitous joint model, strain softening model, and
double yield or cap model. Some of these have plastic yield and possible shear failure;
in addition some tensile failure was governed via simple tensile strength cut-offs
(Lightfoot and Maccelari, 1999).

There are many non linear criteria available in ELFEN; the most widely used in rock
mechanical applications, is the Mohr-Coulomb model combined with elasto-plastic
Rankine (with the option for a rotating crack). More detail on this model is given in
Appendix B.

The Mohr-Coulomb model within ELFEN is applicable when considering failure as a


result of compressive stresses. This is a contrast to the Rankine formulation which
considers only purely tensile stress, and thus allows ELFEN to model fracture
extension from existing discontinuities and ultimately the kinematic release of a failure
block. If it is solely the Mohr-Coulomb failure criterion that is applied ELFEN is unable
to simulate kinematic release; instead analysis reveals yielding within discrete
elements.

With discontinuity controlled failure, a tensile fracture criterion (Rankine) is often


sufficient. When simulating step-path failure/fracture extension through the intact rock
mass, the consideration of failure modes has to be expanded to encompass
compressive fracturing and shear-type failure; this can be done through the
implementation of the Mohr-Coulomb model. The rotating crack fracture model that is
available within ELFEN is based purely on tensile fracture; the combined Mohr
Coulomb with Rankine tensile cut-off model allows both tensile and shear related
fracture (Coggan et al., 2003).

248
APPENDIX A: Important parameters within dynamic fracture-based code

When using a Rankine Rotating Crack criteria within ELFEN, it is necessary to input
values for the Fracture Energy and Fracture toughness, these are defined below.

A.1.1 Fracture Energy


As discussed in Section 2.2.2 the Fracture energy (Gf) controls the ability of fractures to
propagate through the material. Cracks will only propagate once the energy released
per unit crack surface reaches a critical value (Gc), which exceeds the fracture
toughness (KIC) (Whittaker et al., 1992). The fracture energy is related to Young’s
Modulus (E) and the fracture toughness, which should be both known for the material
that is to be used within ELFEN, the Gf can be calculated through the below equation:

G f  IC
2
K
[A.1]
E
see Equation 5.2 for n example along with units.

Normally the area under the ‘softening slope’ is equal to the fracture energy, although
under dynamic conditions this is not the case due to the effects of inertia on the micro-
mechanical response (Yu, 1999 in Owen et al., 2004).

A.1.2 Fracture Toughness


The fracture toughness (KC) is a parameter denoting a condition that is above the crack
tip stress intensity factor (K); thus at situations of stress which are equivalent to the
fracture toughness, crack propagation can occur. The fracture toughness this can be
used as a parameter for rock material classification, or an index for rock fragmentation
processes (Whittaker et al., 1992); with respect to ELFEN it is used as a material
property in the modelling the fragmentation of rock

Critical values of crack tip stress intensity factors exist for all three modes of fracturing,
(KI, KII and KIII), which are described by Whittaker et al. (1992) and Atkinson (1987):
1. mode I where fracturing is tensile related, where the movement of the fracture
due to normal stress, is symmetrical;
2. mode II where movement is due to in-plane shear stress, (mixed mode I-II is
opening and sliding displacements); and
3. mode III where movement is due to anti-plane shear.

249
APPENDIX A: Important parameters within dynamic fracture-based code

Consequently crack initiation can occur when the K, for each of these modes, is at the
critical value KC. For example when the crack tip stress intensity factor (KI) for mode I
fracturing reaches a critical value, called the plane strain fracture toughness (KIC), crack
initiation occurs under mode I loading. This is also the case in mode II, when crack tip
stress intensity factor (KII) reaches the mode II plane strain fracture toughness (KIIC),
crack extension will start (Whittaker et al., 1992). Common situations of fracturing are
under the mixed mode I-II loading. There is no reference to the fracture toughness
concerned with mode III fracturing presumably because this is the most infrequent
action under which crack initiation can occur.

KIC can be calculated from the Gf and E as is shown above as Equation A.1, another
empirical correlation which is proposed by Zhang (2002) is presented below in
Equation A.2:

 t  6.88K IC [A.2]

There are also many laboratory-based methods reported from which a value of KIC can
be determined. The values of KIC listed within Whittaker et al. (1992) range widely from

4.2MPa m for a limestone tested by a method of ‘single edge notched plate in

tension,’ to 0.006 MPa m for a coal tested by a ‘single edge cracked Brazilian disk in
diametral compression. The reader is referred to Potyondy (2002a) for brief detail on
the derivation of fracture toughness from Brazilian strength.

A.2 Discrete Global Properties and surface-type properties


Unlike continuum methods, discontinuum models include parameters such as damping
or contact properties that can be assigned to the discretized elements (Elmo 2006).
Both the discrete global properties and the surface-type properties are labelled discrete
elements constraints used within ELFEN. The discrete element contact data controls
the interaction between the strata within the model (Rockfield, 2008). Liu et al. (2005)
use the below illustration, Figure A.1, to demonstrate the contact laws used in their
DEM approach. As shown in Figure A.1a they use the dashpot analogy, and also refer
to the incorporation of the stiffness of the springs in the calculation of the forces at the
contact. The same contact theory is applied within ELFEN, only in the case of slope
modelling the discrete elements are not disc shaped, instead these contact laws exist
within discontinuities and between discrete domains.

250
APPENDIX A: Important parameters within dynamic fracture-based code

Figure A.1: Contact laws in a shear box DEM model and box description
used during the simulation of a direct shear test: (a) contact model between
nodes; (b) the DEM shear box geometry details (after Liu et al., 2005).

The following sections give an overview of the discrete global properties that are
required within ELFEN, with sections laid out in the order they appear within the Global
properties (of the discrete element constraints) within ELFEN. Each subsection
attempts to explain the particular rationale behind each parameter, justify suggested
values and gives brief examples of their influence.

A.2.1 Damping
A common analogy of damping is a dashpot, this is velocity dependent acting to
slow/deadening or diminish forces. In more detail damping imposes a resisting
influence on oscillations and/or vibrations generated as a result of the sudden
application of force to nodes within the model the model. This sudden application of
forces occurs when the model is loaded or there is an impact of a sliding or falling
block. Therefore it is important to have some kind of damping within the model during
the loading stage and potentially in the failure stage of a model. This damping can be
implemented via the application of point damping, which can be assigned to selected
entities within the model this acts to prevent shock loading and it’s accompanied
vibrational effects.

The viscous damping technique is implemented within ELFEN to achieve element


stabilisation (Hancock, 1973 cited in Rockfield, 2008). In natural systems energy
dissipation is complex, and tends to be proportional to displacements and not velocity.
None the less, damping is required because provided the material remains elastic,
there is no energy loss within a finite element model, and thus the energy wave will
continue to rebound indefinitely if there is no damping (Rockfield, 2006). Damping is
251
APPENDIX A: Important parameters within dynamic fracture-based code

an important method of dissipation that is required due to dynamic formulation of


numerical models, within which steady-state solutions cannot be achieved if energy
dissipation is solely through frictional sliding (Liu et al., 2005).

Hart (1991) describes the two types of damping within UDEC and 3DEC, these are
outlined below:

1. Mass proportional (viscous) damping is the application of a force that is


proportional to the (mass) velocity but in the opposing direction.
2. Stiffness proportional damping, which applies a force that is proportional to the
incremental force or stress; this is normally applied to contacts or stresses in
zones.

Collectively both these types are termed Rayleigh Damping, which is therefore
proportional to both the mass and stiffness of the structure. The combination of mass
and stiffness can be appreciated through the below equation:

C  M K [A.3]

Where M is the mass matrix of a structure, K is the stiffness matrix of a structure, 


and  are constants to be determined from two given damping ratios (Elmo, 2006).

The aim of damping a system is to achieve stability, removing the vibrational effects of
shock loading. Therefore to evaluate the equilibrium that is required within each
element at each time step, the total deformation across the damper must be accurately
calculated.

Four types of damping can be applied within ELFEN: Point (displacement and
rotational) damping, global damping, contact damping and Rayleigh damping; these
are all listed as mechanical constraints within ELFEN. Although contact damping is the
only damping parameter that is assigned within the discrete element properties, the
other forms of damping are also discussed within this subsection to complete the
holistic review of damping within ELFEN.

Point damping (can also be referred to as ‘displacement damping’) has been applied to
the discrete problems considered in models within this research so far. Point damping
is most widely used form of damping, applying velocity-proportional damping to nodes
on selected entities (Rockfield, 2008). The suggestion from Rockfield (2008) is that the

252
APPENDIX A: Important parameters within dynamic fracture-based code

degree of point damping should be between 75 and 100% for dynamic relaxation
simulation of linear problems; lower values should be applied along with appropriate
loading ramps to maintain quasi-static conditions in nonlinear problems. The value
referred to reflects a percentage by which the critical damping factor (automatically
calculated by ELFEN), is multiplied by (Rockfield, 2006). Also Rockfield (2006)
suggest values of 3 to 5% are more advisable to rock slope situations; although they
use 1% within the recent worked example ‘WPSX009’ which involved circular soil
slope.

It is suggestible that a compromise, concerning the loading rate and damping


coefficient, has to be reached during the design of a model. Elmo (2006) implemented
a value of 10% for point damping during the only the primary loading stage, allowing
the system to converge at a steady state prior to the failure stage when point damping
was completely removed. This contrasts to the method that Lobao (2006) employs
which involves point damping throughout all stages at a value of 10%.

Values from this research


Within this research, it has been found that in some cases a point damping of 10%
inhibits failure. Thus lower degrees of point damping were experimented with. This
was to allow an increase in displacement and subsequent failure when surface
conditions dictate the factor of safety (FOS) to be less than one, and no failure when
the FOS is more than one.

These findings correspond to those of Elmo (2006), who stated that both the extension
of existing fractures and the initiation of new fractures are inhibited by high damping
coefficients. This is why Elmo (2006) used a two-stage process with point damping
only present in the first stage, as an attempt to minimise the influence that point
damping would have on the fracturing process. The modelling concerned with this
thesis has followed the staged procedure described in Chapter 3, in some cases
models have damping as only active during the primary loading stage (see model
database).

The application of damping to only the primary loading stage raises an interesting
dilemma to the study of rock slopes, where it has been found that in some cases point
damping is required during the failure stage when blocks may impact on one of the
lines defining the slope surface. If damping is not active during the failure stage, there
is no damping to prevent the impaction generating shock loading and corresponding
dynamic effects (artificial failure). Consequently it in the author’s opinion that if

253
APPENDIX A: Important parameters within dynamic fracture-based code

modelling a slope as a single-surface model, a degree of point damping should be


assigned to the surface for the entirety of the simulation. The actual degree of point
damping that should be assigned requires calibration.

In addition to the above, Elmo (2006) found that the value of point damping that was
assigned did not have such a great influence on simulations with high loading rates.
Elmo (2006) suggested that this is because a higher loading rate introduces dynamic
effects, ultimately leading to a faster velocity of propagation for the cracks.

As well as point damping, global damping is available within ELFEN. Global damping
requires a range of frequencies to be established from which a good control is obtained
of the damping throughout the whole system, unlike point damping which
predominantly damps the lower modes of deformation (Rockfield, 2008). However
there is little guidance on to what kind of problems this should be applied to, and the
description given to the justification of individual frequency requirements within the
explicit ELFEN user manual (Rockfield, 2008) is unclear; consequently this has not
been applied as part of this research.

The final form of damping is contact damping, which is assigned via the global discrete
element constraints window. The viscous contact damping algorithm increases the
specified global normal penalty acting between the discrete elements. The contact
damping may be increased or decreased depending on the velocities (vectors) of the
contacting bodies defined in the analysis; the effect of contact damping increases with
increasing magnitude of the damping factor (Elmo, 2006).

During any numerical simulation contact surfaces frequently alternate between contact
and non-contact states; contact damping modifies the penalty force thus softening
oscillations and minimising the high frequency oscillations (“noise”) that is unavoidably
introduced when the contact surfaces alternate between these states (Rockfield, 2008).
It is advised from Rockfield (2008) that a high value for the contact damping increases
the damping effect and the value for the contact damping parameter is non-
dimensional.

The actual value that one is to apply for the contact damping is again vague, but at
least has a range. Within the ELFEN help file (5.10.3.1 Global Properties) and
Rockfield (2006) it advises that values between 5 and 75%, for general contact
interaction, and the contact damping type set to ‘Velocity-Momentum.’ Contradicting
this is information within the appendix D of explicit ELFEN user manual (Rockfield,

254
APPENDIX A: Important parameters within dynamic fracture-based code

2008), which states that: ‘Contact damping is compulsory, having a ‘maximum of 50


and a min 100%.’ Within the primary simulations of this research contact damping was
set at 30% following the models of Elmo (2006). Although with the advice given at the
meetings with Rockfield on the 11th and 12th July 2006, the contact damping was
subsequently reduced within simulations to 10%; this made no discernable difference
to the simulation results but could be an area that needs more investigation.

A.2.2 Field
The ‘Field’ is a property within the discrete global constraints that defines the thickness
of the contact layer and the subsequent maximum penetration which is permitted. This
is used within the second (local) part of the contact search process that is employed
within ELFEN (basis of information is from Rockfield (2006)):

1. Firstly a list of nodes close by is generated via a global search of nodes that fall
within a distance that is less than the Zone (see Section A.2.4); this generates
a ‘candidate list;’
2. Following this a local search is performed upon this candidate list, at every
timestep; this establishes which elements are in contact and accordingly
evaluates their contact forces using the penalty numbers and calculated
penetration (overlap) within the basic contact force equation. The more
complex the force equation that is actually used within ELFEN, entails
implementation of the other parameters that are detailed within this Section.

The field relates to a function of the length of the element size that is selected by the
user; the suggested range is between 10-20% of the smallest element size in the mesh
(Rockfield, 2008). This value applied can be adjusted, if it is found that from running
the analysis one or more nodes penetrate though an element; however it is advised
that the field should not exceed 20% of the size of the finite element (Rockfield, 2008).
If the overlap (penetration of an element into another) exceeds the Field, then contact
will go undetected and presumably the penetrating element will pass straight through
the element it should have made contact with.

If a large field is assigned then the time step should be small (Rockfield, 2008);
consequently it should be presumed that with a large field the contact layer is thick.
Thus to reduce the penetration the time step should be small to decrease the likelihood
of a change in acceleration, displacement or in this case penetration, during the
duration of the time step. Consequently penetration of a node through an element may
occur when the field size is set at a value that is too small for the model in question;
255
APPENDIX A: Important parameters within dynamic fracture-based code

this corresponds with the information in Rockfield (2008) that states that if the field is
too low penetration occurs because elements will have moved more than the field
length within a single timestep.

A.2.3 Penalties and Stiffness


The contact and global penalties within ELFEN are contact forces between nodes and
facets within the numerical model that prevent penetration of the surfaces. The
penalties coefficient is a stiffness term with the units N/m, although due to the
convenience of working with contact stresses as opposed to contact forces, the penalty
coefficient is considered a modulus relating stress and displacement with the unit Pa/m
(Klerck, 2000).

Klerck (2000) relates contact stiffness to a point of contact between two springs of
known normal and tangential stiffness, (  n and  t respectively), as shown in

Figure A.2. ELFEN defines the penalty contacting couple as a modulus relating stress
and displacement, which is calculated from the stiffness of each of the springs,  n and

 t (Elmo, 2006). Within ELFEN the  n and  t of each of these springs is

represented by the normal penalty and tangential penalty respectively; this research
follows the notation used by Elmo (2006), which was to indicate the normal penalty
coefficient by (Pn) and tangential penalty coefficient by (Pt). The penalties need to be
set with correct values; if too low the nodes can penetrate one another and if too high
then the material can behave in a brittle manner as Elmo (2006) found.

Figure A.2: The penalty contacting couple in ELFEN


as an equivalent spring system (after Klerck, 2000).

Elmo (2006) modified the information that Pine (2006, personal communication) gave
forming the below illustration. As well as allowing the derivation of joint normal
stiffness, shown in Equation A.6, Figure A.3 can be also used to describe the fact that
256
APPENDIX A: Important parameters within dynamic fracture-based code

lower penalty equates to a decreased degree of freedom for the discretized blocks;
consequently assignment of a lower penalty permits rotation.

Figure A.3: Explanation, using a series of small pillars, of the normal penalty (Pn) in terms
of the intact material Young’s modulus (E) from the information that Pine (2006, personal
communication) gave (after Elmo, 2006).

Formulae that are associated with Figure A.3 are presented below in equations A.4 to
A.6:

Strength of each pillar (σp):  p  


r2
2
[A.4]
wp

p h p  hp r2
Displacement measured for each pillar (δp):  p   2
[A.5]
E E wp

  
Joint normal stiffness  : 
2

  
E wp
 p 
[A.6]
p hp r 2

As the penalties are closely related to stiffness, normal and shear stiffness has been
researched. The joint stiffness parameters are fundamental properties in the numerical
modelling of jointed rock, describing the stress-deformation characteristics of a joint
(Wines and Lilly, 2003). An example of a variable application of stiffness is in the DDM
approach, which is often applied to the modelling of a tabular ore body; to simulate
mining the material stiffness is reduced to zero within mined out areas inside the
modelling area (Elmo, 2006).

In the normal direction, the stiffness governs the relationship between stress and
displacement (Kimber et al., 1997). Klerck (2000) equates normal stiffness to surface
irregularities such as asperities or joint gouge/filling. The stiffness within intact material
and on joints is related to stress and displacement; the stiffness acting in the normal
257
APPENDIX A: Important parameters within dynamic fracture-based code

direction, is assumed to be linear (Hart, 1991), following the formula shown in Equation
A.7.

 n  k n u n [A.7]

where  n is the effective normal stress increment,


u is the normal/displacement increment.

Depending on the joint spacing the normal stiffness of joints is considerably lower than
that of the intact rock separating the joints (Barton, 1972). The joint normal stiffness is
influenced by the below factors:

1. The initial actual contact area;


2. The thickness, type and physical properties of any infill material;
3. The strength and deformability of the asperities, and
4. The joint wall roughness (Bandis et al., 1983 in Wines and Lilly, 2003).

The joint normal stiffness can now be estimated via appropriate laboratory testing. At
one time plate jacking tests were one of the most frequent forms of rock mechanics
testing; these were often conducted to obtain a deformation modulus for rock mass
within dam sites, for comparison against the E of the concrete dam.

In the shear direction, the shear stiffness (Ks) is constant and controls the cross-joint
response (Kimber et al., 1997). The shear stiffness is reduced to zero when sliding
occurs as a result of interfacial stresses exceeding the frictional resistance (Huang et
al., 2002). Wines and Lilly, 2003 present the following equation for the estimation of
the peak shear stiffness (MPa/m), which Barton and Choubey suggested in 1977:

[A.8]
where r = residual frictional angle,
JRC = Joint Roughness coefficient,
Lx = the joint length in metres

Wines and Lilly (2003) found that joint shear stiffness calculation via Equation A.8,
provided significantly lower results than the range of joint shear stiffness that were
calculated from direct shear testing. This could have been due to scale effect because

258
APPENDIX A: Important parameters within dynamic fracture-based code

direct shear testing was conducted on samples with significantly smaller than mean
trace lengths (Wines and Lilly, 2003).

The value of penalty that is assigned within ELFEN is dependent on the contact
pressure and material properties of the objects which make contact with each other
(Rockfield, 2008). The range that the Pn must fall within is 0.5E to 2.0E, given by
Rockfield (2008); although typically the Pn is given a value that is approximately equal
to the average E of the material. The advice given within the ELFEN help file
(Rockfield, 2008) is that the tangential penalty should be about 0.1 of the normal
penalty.

During a training course with Rockfield, it was stated that normal penalty governs the
contact force that is generated between elements and therefore determines the amount
of penetration (overlap) between elements and the amount of subsequent energy that
is introduced into the system. If the Pn is set too high then the resulting contact forces
that will be generated will be high and consequent violent rebounding and extreme
distortion of elements can occur; if too low than contact forces will be insufficient to
repel penetrating elements and thus objects will be able to pass through each other
(Rockfield, 2006). In addition to this it was stated that a large Pn will decrease the
timestep as they have the following relationship:

t cr 
Pn
M1  M 2
[A.9]

(where M = mass of element)

Once a simulation has been run, if it is found that one or more nodes penetrate though
an element, Rockfield (2008) suggest that the simulation should be re-run with either
an adjustment made to the field, or the penalty term should be increased. In addition to
this a reduction in the factor of critical time step may be required in order to detect the
contact sooner (Rockfield, 2008).

With respect to fracturing, Elmo (2006) found that that when the penalty values are
high the pre peak degree of fracturing is negligible, but at post peak the degree of
fracturing is higher and explosive like; thus proving the fact that there is a higher joint
stiffness at high penalties. Elmo (2006) also found that locking-up effects are
proportional to penalty values, with a brittle response resulting from the use of high
penalties and rotation and translation occurs with low penalties. Finally Elmo (2006)
additionally experimented with the Pn to Pt ratio; at low ratios the response of the

259
APPENDIX A: Important parameters within dynamic fracture-based code

modelled jointed material was stiffer and stronger. In summary Elmo (2006) found that:
‘At high Pn values the response of the pillars was more brittle and thus the estimated
strength of the pillars would be lower.’

A.2.4 Zone
This is a parameter that is defined within the discrete element constraints. Within the
ELFEN user manual (Rockfield, 2008) it is referred to as the ‘buffer zone;’ with the
purpose of defining the area within which local nodes are searched during contact
detection. Therefore it is used to locate nodes close to any surface during contact
detection (Rockfield, 2008); it is also the first part of the two-level search process
(described in Section A.2.2), that is employed within ELFEN to determine which bodies
are interacting. The advice given within the user manual is that the zone should be
given a value that represents the average size of the side lengths of the finite element
mesh that is assigned to the problem. During the ELFEN training course (Rockfield,
2006), it was advised that the zone should be set at small values within slope problems
as velocities should be slow.

A.2.5 Smallest Element


As the name of this constraint suggests, this should be set as the size of the smallest
element that will be created when using a fracturing analysis, which is defined when
assigning the non-linear criterion. This can be assigned a value that is larger than the
smallest element if one desires no fracturing within the elements. Alternatively if it is
set at a value smaller that than the actual smallest element, fracture propagation can
occur through the meshed elements rather than being limited to fracturing along
element boundaries, as would occur if it was correctly set at the actual smallest
element size.

However, as suggested by Rockfield (2006), if the smallest element is not set within a
model where fracturing is set to occur, then elements could in theory continue to
fracture into extremely tiny elements; they consider this to be unrealistic and it will also
result in a unnecessarily small timestep. Perhaps a better view of the smallest element
constraint value is that it is a control on what smallest element the simulation needs to
produce in order to link fractures. Although further guidance within Rockfield (2006)
suggests that setting the smallest element so that fracturing can occur through the
meshed elements, can slow the simulation runtime considerably as very small
timesteps can occur as a result of ‘badly shaped’ elements. Perhaps as well as the

260
APPENDIX A: Important parameters within dynamic fracture-based code

control on the smallest element that is formed from the simulation, there should be a
control on the shape of element; either this or the mesh size needs to be set a small as
is feasible within domains where fracture propagation is desired.

A.2.6 Contact Damping Type


There is a limited degree of guidance within the ELFEN help file or the ELFEN user
manual (Rockfield, 2008), on the applications of the different types of contact damping
that are available within ELFEN. There are four different types of contact damping
available within ELFEN, the default contact damping type is Velocity\Momentum; this
has been applied within all models concerned with this research. Within Rockfield
(2006) Velocity\Momentum is recommended for general mechanical interactions;
information on the other types of contact damping is sparse although within the ELFEN
help file (Rockfield, 2008), a viscous contact damping is recommended for simulations
with particulate geometry.

A.2.7 Contact Type


There are two types of contact that are available through the ELFEN 3.8.7 interface,
node-to-edge and edge-to-edge. There is one contact-type that is not on official
release yet and subsequently has to be applied via modification of the neutral file
(facet-to-facet contact algorithm); this is discussed within Section A.5.1.

The default contact type is node-to-edge, which detects contacts to be between the
nodes of one surface and the element edges on the other; edge-to-edge contact type is
also available but uses only the element edges to detect contact (Rockfield, 2008).
Simulations of 3D geometries and adaptive projects may benefit from the application of
the node-to-edge contact type, although an increase in the penalty numbers may be
required to compensate for the increase in contact detection zone (Rockfield, 2008).

Following the meetings with Rockfield on 11th and 12th July 2006, it was suggested that
the facet-to-facet contact can also now be applied as opposed to the contact algorithms
that are discussed above; this is described in Section A.5.

261
APPENDIX A: Important parameters within dynamic fracture-based code

A.3 Factor of Critical Step


The factor of critical step is one of the four time step controls (Initial, Maximum,
Maximum Change and factor of critical step), defined within the analysis control of
dynamic (explicit and discrete) simulations. The factor of critical step is closely related
to the ‘critical timestep’ (∆tcr) within the solution algorithm, it is used to multiple the
current value of the ∆tcr (which is re-evaluated at every stage), to obtain the actual
working timestep (Rockfield, 2006). The ∆tcr relates to maximum timestep permissible
during the simulation, and it is based on the smallest element because of the central
difference formulation that occurs within ELFEN (i.e. the stress wave generated within
the model must only pass through only one element during each timestep; if during the
first timestep the stress wave has already passed through the smallest element then
the solution becomes unstable).

During each timestep the contact at each of the elements within the model is
evaluated. Advice given in Rockfield (2006) suggests that if the ∆tcr is assigned as a
small value, the simulation is forced to have small timesteps. This is particularly
important within models where there maybe high velocities/high speed impact; in such
cases it is possible for elements to overlap (penetrate), during a single timestep, further
than the distance that is set in the Field (see Section A.2.2). During such an event
contact will not be detected and the elements will pass straight through each other
(Rockfield, 2006).

Kimber et al. (1997) notes that the time-stepping algorithm in UDEC allows the
program to model dynamic behaviour; presumably this is the same in all DEM
techniques that utilise a time-stepping algorithm. During each time-step in a UDEC
simulation, two calculations are performed, firstly the velocity and direction of each
element is calculated via the application of Newton’s second law, and secondly the
contact force across every discontinuity in the rock mass is established through the use
of force displacement law (Kimber et al., 1997). All Itasca software appears to uses
this time-stepping procedure:

 Lightfoot and Maccelari (1999) note that FLAC recalculates the forces,
accelerations and velocities being at each grid point, with each time step;
subsequent stresses within the continuum can be derived from strains
calculated from grid point movement.
 Potyondy (2002b) notes that the time step within PFC is adjusted automatically
according to local conditions and therefore the simulation is fully dynamic.

262
APPENDIX A: Important parameters within dynamic fracture-based code

Rockfield (2008) suggest that to ensure that the simulation remains stable, the factor of
critical step should be about 0.9 for 2D and 0.7 for 3D; a lower value is also advised for
particular simulations such as high speed impactions. Rockfield (2006) suggest a
factor of critical step of around 0.25 for dynamic problems or simulations where
fracturing is likely. If a very large initial time step is assigned to the model then ELFEN
will automatically calculate the ∆tcr (Rockfield, 2006). Klerck (2000) notes that a small
time step also allows the model to act more like an elastic continuum consequently
reducing penetration between the blocks being modelled.

Rockfield (2006) present the following derivation for ∆tcr:

Δt c r  where c 
l E

[A.10]
c

of the system), E is the material intact Young’s modulus and  the material density].
[ l is the length of the smallest mesh element, c is the wave speed (highest frequency

Consequently the ∆tcr increases with increasing mesh/smallest element sizes,


increasing density of the medium being modelled and decreasing material stiffness
(Rockfield, 2006). Therefore the ∆tcr should be set at a value that relates to the
smallest element denoted within the mesh, the E and the  of the intact material in
question.

Elmo (2006) set a ∆tcr of 0.75 for the loading stage and 0.1 for the second failure stage,
any longer than this then the processing requirements for each model would be
increased, although this is a compromise as a too shorter ∆tcr can have adverse effects
on the explicit solution convergence. In addition to this Elmo (2006) also found that the
∆tcr is directly proportional to the minimum mesh size. Klerck (2000) discovered that
displacements for certain problems increased with increasing time steps; he changed
the time step between 0.1e-5 and 0.1e-4.

A.4 Mesh Assignment


This section attempts to give some guidance on the mesh assignment within ELFEN,
describing the definitions and settings that have been used within this research. In
particular the following section, describes the influence of mesh density/element size
on the simulation outcome, attempting to give some guidance on the appropriate mesh
size to apply to design problems.

263
APPENDIX A: Important parameters within dynamic fracture-based code

The mesh assignment within ELFEN is complex and involves the application of a mesh
generation type, element type, and mesh density definition. Throughout this research
the ‘unstructured’ mesh generation method has been used, there is also a structured
method which requires the 2D geometry to be split up into four-sided surfaces
(Rockfield, 2006). Instead the unstructured mesh creates triangular or quadrilateral
elements in 2D models, and tetrahedral elements in 3D models.

The two generation algorithms available in ELFEN when generating an unstructured


mesh, are ‘Advancing front’ and ‘Delaunay’; in all models within this research, except
those simulations concerned with Delabole, a Delaunay generation algorithm has been
used. The Delaunay method was chosen following the example WPSX007 within
Section 3.3.7 of the ELFEN worked example manual, which dealt with the excavation
of an underlying coal seam (Rockfield, 2008), and also the work of
Elmo (2006) whom applied the Delaunay method. A recent contradiction to this is the
advice given during the ELFEN training (Rockfield, 2006) when it was stated that slope
models should use the advancing front algorithm, as was used within the Slope
Example WPSX009, (Rockfield, 2008).

It was also stated during the ELFEN training (Rockfield, 2006) that slope models
should use an unstructured mesh. During the ELFEN training (Rockfield, 2006) it was
suggested that: ‘The Advancing Front mesh algorithm generally produces a smoother
mesh.’ This could therefore potentially limit fracture propagation. Instead it was
suggested that the Delaunay algorithm produces a more random mesh and this could
be more applicable to slope problems where fractures should not be so straight (limited
to the boundaries defined within the mesh), as could occur with the Advancing Front
algorithm.

A.4.1 Mesh Density


The mesh density, also known as the element size or mesh size, defines the number of
elements within any given section of the domain; this is often critical in generating
accurate results from the finite element model (Rockfield, 2008). Elmo (2006) states
that potentially the mesh size chosen can have a greater influence on the mechanical
behaviour of a model than the strength parameters of the intact material; although that
relationship does become less prevalent when a discrete fracture network (DFN) is
incorporated within the model.

264
APPENDIX A: Important parameters within dynamic fracture-based code

There is very little guidance on the actual mesh size that should be assigned to
problems; the only suggestion from Rockfield (2008) is that a higher mesh density is
required in regions where sharp changes in stress occur. The mesh density chosen
depends on the scale on which displacements and stresses are being examined and
the processing speed that is available; for instance in larger scale problems which
examine slopes of tens of meters in height, the mesh size should not be less than at
least 2m. If fracture propagation within a certain part of that slope is being examined,
then a finer mesh density can be assigned to the area surrounding the feature.

Throughout most of the models concerned with this research, a relatively coarse mesh
size has been used; also importantly it has been fixed (un-graded). Barla et al. (2003)
used a graduated mesh within their continuum model comprising 25,254 elements.
Models with a small mesh size and subsequent large number of elements generally
give a better approximation to the real situation (Curran and Ofoegbu, 1993). Elmo
(2006) ran his 3D uniaxial models with element sizes one tenth of the model width;
however some of these took in excess of ten days to run.

A large mesh size within 3D models controls the hardening behaviour by preventing
shear bands developing within the 3D space. The hardening response is a subsequent
result of 3D blocks remaining in place as opposed to being kinematically released once
fracturing occurs; this allows a greater confinement than within similar models with
smaller mesh sizes (Elmo, 2006). In contrast to this Elmo (2006) notes that with
models with a small mesh size, the result may be shear band formation as a
consequence of the coalescence of abundant parallel tensile fractures.

It is important to note that the mesh size definition is particularly complex within ELFEN
when there are small spaces between tight and interlocking fractures, which can be
generated from FracMan modelling. A consequence of these closely spaced fractures
is that their representation has to be through small tetrahedral elements which can lead
to significantly increased runtimes (Elmo, 2006). Therefore the fracture network one is
modelling has to be considered carefully, with regard to the ability that the present
single processor version of ELFEN has to model the behaviour between fractures. In
the same way, the mesh density selected within ELFEN has to be applied with respect
to the spacing of fractures within the material that is being modelled; one possibility is
to mesh the problem with a finer mesh surrounding areas of interest or closely spaced
fractures, and a coarse mesh where movements are presumed to be minimal.

265
APPENDIX A: Important parameters within dynamic fracture-based code

A.5 Additional edits to the neutral file


During the progress of this research, it has been discovered that in some cases
particular edits are required within the neutral file. Some advice is given in the
following sections on the action of some of these particular edits, and more importantly
to which models they are appropriate.

A.5.1 Option 166


It has been suggested by both Lobao (2006, personal communication) and Rockfield
(2006) that the facet-to-facet contact algorithm (‘option 166’) can be beneficial in slope
problems, as substitute to the other contact algorithms that are discussed in Section
A.2.7. From the research of this thesis, it appears that the main advantage is that it
can supplement the release of failure blocks, as reported in Section 6.2 (in particular
Figure 6.6c).

Rockfield (2006) also suggested that the facet-to-facet contact algorithm is beneficial
because in the case of penetration it takes into account how much area of another
element penetrates the element in question and also where it has come from.
Presumably this then aids in the more accurate direction of the discrete elements.
However, the facet-to-facet contact algorithm is particularly advantageous in 3D
models, and might not make a difference within 2D models (Rockfield, 2006).
Therefore the benefits of the contact algorithm may not be apparent within this thesis,
due to the lack of 3D modelling.

As referred to in Section A.2.7, the facet-to-facet contact algorithm is not on official


release. Therefore to enable it, the neutral file must be edited. Option 166 must be
inserted under the ‘system’ heading (as indicated in Figure A.4).

System {
Options { 1
166
}
}

Figure A.4: Coding that must be inserted into the


neutral file to ensure the facet-to-facet contact
algorithm is used.

266
APPENDIX A: Important parameters within dynamic fracture-based code

Influence of Option 166 in the models of this research


In addition to the finding noted above, where the facet-to-facet contact algorithm aided
in the separation of a failure block, it was also found that within a planar failure slope
model the vertical displacement of a failure block decreased. The data supporting this
is presented in Appendix H (Figure H.3). Further testing is necessary to verify the
influence of the facet-facet contact algorithm in slope models.

During shear box simulations (which are not reported as part of this thesis), several
aspects changed when using the facet-to-facet contact algorithm:

1. Stress concentrations within the centre of the shear box model were prevented.
2. The upper shear block was stopped from being completely separated following
100% shear pressure application.
3. Shear stress was slightly increased although the shear stress-shear
displacement curve became truncated at the elastic-plastic boundary with
models ran in ELFEN version 3.8.3.

A brief re-run of the shear box simulations was conducted using a newer version of
ELFEN (3.8.5). It was discovered the implementation of the facet-to-facet contact
algorithm made no difference to the stress distribution, as even within a model ran
without option 166, there was no stress concentration within the centre of the model.
However, without option 166 the contact pressure is significantly lower than within a
model with option 166.

A.5.2 Option 37
This is an edit that is available through the present ELFEN interface, controlling an
algorithm that is of particular use in models where plasticity is expected to occur.
Within the pre-processor this can be enabled by activating the element option 37
(‘tetrahedral element’), within the control window.

When enabled, an averaging algorithm is used, which improves the stress distribution
across elements; in particular Rockfield (2007, personal communication) suggest that
the algorithm can partly compensate for the over-stiffness of triangular elements.
Section 5.2.4 reports the influence of the tetrahedral element control, providing
evidence that further development of strain and consequent deformation only occurred
when option 37 was enabled.

267
APPENDIX A: Important parameters within dynamic fracture-based code

It is recommended that there should be further research into the influence option 37.
Rockfield (2007, personal communication) state that the algorithm is still under
development, and there are currently some problems related to using the tetrahedral
element control with pressure-sensitive material models, (giving the example of Mohr-
Coulomb criterion). Problems with the tetrahedral element control have not been
encountered directly during this research, and therefore it is recommended that if shear
strain and deformation are not occurring within a model where it is expected, a
simulation with an active tetrahedral element control should be run. During this
research this was only encountered during the simulation of shear strain development
within a chalk cliff, as reported in Section 5.2.4.

A.5.3 Option 179


It is stated in the training manual (Rockfield, 2006) that ‘Option 179’ is necessary with
the constraint release method, so that the fixed sections of the slope can be identified
for relaxation. This is only necessary with the structural-constraint-release approach
and not the applied displacement-constraint-release method.

Detail on the methods of slope release is given in Section 3.1.1. In particular during
the research of this thesis, it was found that a structural-constraint approach using a
structural-fixity is not possible in a slope model with embedded discontinuities.
Consequently a structural-constraint approach cannot be used in discontinuous slope
models. Further information on the action of option 179, and detail on how to remove
the beam element, is required prior to future structural-constraint-release modelling.
Instead, during the modelling of this thesis, an applied-displacement approach has
been developed as an alternative (see Section 3.1.1), removing the need to implement
option 179.

268
APPENDIX B: Rock mass criterion used in ELFEN

APPENDIX B: Detail on rock mass criterion used within


ELFEN (Mohr-Coulomb with Rankine tensile cut-off)

The Mohr-Rankine criterion is one of the compressible elastoplastic material models


that are available within ELFEN. The rotating crack fracture model that is available
within ELFEN is based purely on tensile fracture; the combined Mohr Coulomb with
Rankine tensile cut-off model allows both tensile and shear related fracture (Coggan et
al., 2003).

As outlined in Section 2.2.4, the Mohr-Coulomb with Rankine tensile cut-off enables the
simulation of mixed mode I-II failure. Mode I fracture extension is simulated by means
of the Rankine (with fracture) non-linear criterion; whereas mode II yielding, creates
shear bands via the Mohr-Coulomb non linear criterion. However, fracture extension
due to shear cannot be simulated within the current version of ELFEN, and therefore
kinematic release is limited to slope models where mode I fracture extension is likely to
develop a release plane.

Alone, the Mohr-Coulomb and Rankine rotating crack criteria have benefits and
disadvantages. In particular yielding due to shear within a model will not occur when a
purely Rankine criterion is used. The Mohr-Coulomb criterion is empirical and
assumes that the rock will fail in shear; also experimental peak strength envelopes are
generally non-linear and therefore the representation using a linear criterion can be
somewhat of a simplification. These disadvantages to the Mohr-Coulomb failure
criterion give justification to the fact that the Mohr-Coulomb criteria is not ideal for intact
rock, and instead more applicable to the description of the shear strength of
discontinuities within fractured rock (Lightfoot & Maccelari, 1999).

For specific detail on the tensile fracture model within ELFEN, see Elmo (2006).
Importantly once an element has yielded, ‘anisotropic damage evolution’ is represented
by the rotating crack criterion, which down-grades the elastic modulus in the direction
of the major principal stress invariant (Cai and Kaiser, 2004).

Within the Mohr-Rankine criterion the conventional Mohr-Coulomb hydrostatic cut-off is


replaced with a Rankine tensile corner as Klerck (2000) illustrates, (reproduced in
Figure B.1). The grey area highlighted in Figure B.1 represents the region inside of the
failure envelope formed by the combined Mohr-Coulomb with Rankine criteria.
Elasticity is encountered in stress states which fall within the domain beneath the linear

269
APPENDIX B: Rock mass criterion used in ELFEN

yield envelopes (FOS >1) whereas plasticity and resulting failure occurs when stresses
are at values outside of the envelope (FOS <1).

Figure B.1: Presentation of the cut-offs that can be used in different


criteria; the grey area represents the region inside of the failure envelope
formed by the combined Mohr-Coulomb with Rankine criteria (modified
after Klerck, 2000).

Note that as shown in Figure B.1, the convention in ELFEN is for negative stress to
indicate compression, whereas positive stress is tensile; this is due to the original
development of the code as a tool for civil engineers. This contradicts the vice-versa
convention within rock mechanics. Consequently throughout this thesis, the negative
stresses (compressive) from ELFEN, have been transposed to positive stresses, and
the same with the positive (tensile) stress to negative stress. This keeps the
convention used in rock mechanics, so that any stresses outputted from ELFEN are
comparable to other geomechanical software packages.

As outlined in Appendix A, the neutral file is the way in which control is mainly achieved
within an ELFEN simulation. Currently the Mohr-Coulomb with Rankine tensile cut-off
criterion cannot be enabled through the software interface (pre-processor); the easiest
way to ensure that the model is used during the simulation, is to insert the material as a
purely Rankine Rotating Crack material. By doing this the fracturing flags in all stages
are activated which allows fracturing to take place. Importantly the neutral file will then
need editing so that the coding presented in Figure B.2 is inserted in place of the
Rankine Rotating Crack material control.

270
APPENDIX B: Rock mass criterion used in ELFEN

Importantly the ’19,’ inserted into the plastic material flags will enable the Mohr-
Coulomb with Rankine tensile cut-off criterion (“Model 19”).

Material_data { 1
Material_name {
"chalk"
}
Elastic_material_flags { NFGELA { 4 }
0100


}
# E (Pa) ρ (kg/m3) R
Elastic_properties { NMPRP { 15 }
1e+009 0 0 0.24 0 0 0 0 0 0 0 0 1700 0 0
}
Plastic_material_flags { NMFPLS { 2 }
0 19

# c(Pa)  (°) i (°) σt (Pa) Gf (J/m2)


}

Plastic_properties { NPRPLS { 5 }
26400 49.9 5 13000 47.5
}
Failure_material_flags { 3
0 1 1
}
Number_state_variables {
12
}
}
Figure B.2: Coding required for the Mohr-Coulomb with Rankine tensile cut-off criterion, with red
symbols inserted as a reference to what each of the numbers represents (and the units).

271
Appendix C: Analysis methods in SLIDE

APPENDIX C: Slice equilibrium analysis methods within


SLIDE

There are ten, slice equilibrium analysis methods that are available within SLIDE
(Rocscience, 2008). These are Bishop simplified, Corps of Engineers,
GLE/Morgenstern-Price, Janbu simplified, Janbu corrected, Lowe-Karafiath,
Ordinary/Fellenius and Spencer. The reader is referred to Rocscience (2002) and
Diederichs et al. (2007) for a brief overview on the detail of these formulations;
primarily the difference between these methods is the assumptions on which shear and
normal interslice forces are based.

The majority of these methods can be used during the study and resulting FOS
derivation of circular and non-circular failures. The exceptions are:

 The Ordinary/Fellenius analysis is only valid for circular slide surfaces


(Geostru, 2006);
 In addition there does seem to be some question over whether the Bishop
method of analysis is applicable to non-circular failures, as it was originally
developed, like most of the criteria, for analysis of circular shaped failure
surfaces. Rocscience (2002) suggest that the Bishop method is only
applicable to circular failures; however, as reported in Section 5.2.2, it was
found that the Bishop analysis gave a reasonable estimate of the strength of a
non-circular failure.
 Geostru (2006); suggest that the Morgenstern-Price can be considered as the
most popular for non-circular failure as a result of it satisfying the complete
equilibrium conditions with the least numerical difficulties.
 The Janbu method can analyse both circular and non-circular failure. However
Wyllie and Mah (2004) suggest that the application of the Janbu method should
be restricted to shallow slide surfaces where the angle of internal friction () is
>30°, as serious error is associated with the factor of safety analysis of deep
slide surfaces in materials with a low .

Diederichs et al. (2007) reported that the Bishop analysis gave the most realistic limit
equilibrium model when residual strength was used; with comparisons to a Finite
Element method Shear Strength Reduction (FEM-SSR) analysis, within a slope
problem where there are multiple materials.

272
APPENDI X D: Socio-economic factors

APPENDIX D: Socio-economic factors that may


compromise slope stability
During the review of failure within fractured slopes, it has been clear that there are
human-based factors that can act to compromise slope stability. This appendix
contains a brief review of some human based factors that have come to light recently,
which may influence slope stability.

With the ever increasing demands on mining and profit margins, mining practices can
be considered to becoming more aggressive with steeper slopes and greater depths
becoming achievable. The lack of information on existing slopes of such scales is a
significant issue though, with what Robertson (2007) referred to as “Optimistically
aggressive and plain poorly designed slopes,” as a result of inadequate understanding
of rock mass, which can lead to poor geotechnical models in the early design stages.

Not just are there inadequacies in the understanding of the rock mass at the scales that
modern society is demanding, it can also be suggested that there are some poor slope
practices due to inconsistencies in geotechnical approaches. At the Slope Monitoring
Forum of the 2007 International Symposium on Rock Slope Stability in Open Pit Mining
and Civil Engineering, it was recognised that there is a high turn over of engineering
based employees within mining companies. With the increasing mining demands and
a lack of university courses training new staff, employees can find they can change job
and location freely and frequently. This can lead to a lack of experience and skills-
base within certain mines; without the empirical knowledge of the rock mass behaviour
at certain mines, consultation with specialists is required for design-based decisions.

The present state in mining operations is that there is a wealth of data, most of which is
from blanket monitoring schemes; however the data is often poorly managed within
user-unfriendly databases, thus data can often get forgotten. With limited numbers of
staff, engineering based employees can spend a vast majority of their time collecting
data with no time for analysis in-house; instead the analysis is often done by external
consultancies. During a discussion at the 2007 Slope Monitoring Forum, it was
suggested that the aspects mentioned above, could be one of the many reasons to
why there is such a high turn-over of staff within modern mining companies.

It is important that such issues are addressed if large slopes are to be designed with
confidence. It is even in the interest of mine fanciers to consider such issues as slope
failures can cost; Robertson (2007) referred to a single slope failure incident that cost
$38M in 1989!

273
APPENDIX E: Empirical classification schemes

APPENDIX E: Empirical-based rock mass


characterisation schemes

Edelbro (2003) listed the below systems and criteria of rock mass strength derivation:
 Rock Mass Rating (RMR),
 Rock Mass Strength (RMS),
 Mining Rock Mass rating (MRMR),
 Rock Mass Quality (Q-system),
 Rock Mass Number (N),
 Rock Mass index (RMI)
 Geological Strength Index (GSI),
 Yudhbir, Sheorey and Hoek-Brown strength criteria.

In addition to these there are numerous systems that have been developed for direct
application to rock slope strength, such as the Slope Mass Rating (developed by
Romana in 1985), the Slope Rock Mass Rating (developed by Robertson in 1988) and
the Chinese system for Slope Mass Rating (developed by Chen in 1995). The more
favoured systems within this list are the RMR, Q-system and GSI coupled with Hoek-
Brown criterion. Feng and Hudson (2004) suggest that the success of rock mass
characterisation schemes such as the Q and RMR, is a result of their ability to correlate
many rock mechanical properties by taking a set of key parameters into consideration
easily and quickly.

Importantly all of the systems listed above require a combination of an initial


characterisation of the rock mass and subsequent adjustment factors. The
characterisation of the rock mass typically involves factoring to allow for structure,
blasting and weathering culminating in Bieniawski's RMR or a close derivative of this;
this is then adjusted to account for factors such as discontinuity orientation, excavation
method, weathering, induced stresses and major planes of weakness (Duran and
Douglas, 2000).

Generally classification-based methods arrive at a rating (typically out of 100) which


describes the quality of the rock mass, from which slope design decisions can be
made, with the potential of sensitivity studies through altering the respective slope
parameters within the system.

274
APPENDIX F: Excavation ELFEN models

APPENDIX F: A stress-related issue in ELFEN


excavation models

As discussed in Chapters 4 and 5, it has been noted that in order to get kinematic
release of a discrete block within an ELFEN model, any discontinuities which intersect
the slope boundaries need to be extended through into the either the open space or the
excavation block that is to be removed. This can cause problems with stress
equalisation upon discontinuities when applying gravity within excavation models, as
shown in the Figure F.1, which illustrates a concentration of stress around the tip of the
extruding discontinuity.

Figure F.1: Discrete (un-averaged) direct stress y-y (Pa) plot of vertical stress in a simple planar
failure model which is unloaded via an excavation; the screen shot is taken at 100% gravity with
contact initialisation still active. Note the stress concentration around the extruding base of the
failure plane.

This stress concentration occurs whether the model is ramp or drop loaded and
whether contact initialisation is activated or not. The only ways to avoid this stress
concentration is to:

275
APPENDIX F: Excavation ELFEN models

1. Unload the model via the constraint release method as opposed to


excavating the slope;
2. Keep the geostatic initialisation active until the excavation block has been/is
started to be removed;
3. Make the mesh finer around the tip of the extruding failure plane;
4. Or finally increase the time over which loading occurs.

The final two solutions (3 and 4) were suggested by Rockfield (2007, personal
communication). However both these methods would lead to an increased runtimes,
which is undesirable in such a simple model, but possibly necessary if other
simulations where an excavation method is strictly required. The second technique
may be suitable if the slope failure mechanism is to develop as a result of the release
of discontinuities. However if the slope failure trigger mechanism is to be either the
removal of material or a rise in the phreatic surface, then the geostatic initialisation
should occur prior to excavation.

Another issue related to extending the fracture tip through the slope face, is the
decreased time-step and consequent increased runtime due to badly shaped elements
created within the excavation block, between the surface and failure plane tip. This is
particularly an issue where discontinuities are steeply inclined, as the subsequent
angle between the fracture tip and the slope face is very acute, decreasing the quality
of the mesh.

In the case of the simple small-scale model illustrated in Figure F.1, the stress
concentration at the base of the discontinuity was not a problem. However within a
simulation of large-scale slope, such concentrations could become a significant issue; if
high enough, the stress concentration could potentially induce artificial fracturing at the
joint termination within the excavation block, which could extend into the toe of the
slope.

It is suggested that a solution should be sought within Rockfield, to ensure that the
embedded discontinuities that intersect the face do not have to pass through the slope
face. Consequently excavation models would be able to be run seamlessly, as
constraint simulations currently can.

276
APPENDIX G: Additional results for ELFEN planar failure models

APPENDIX G – Additional results for ELFEN planar


failure models

G.1 Data from damping study (slope 1)


At first data was taken from 1 s runtime (as in Figure 4.2a), however it was
subsequently discovered that displacement upon frictional discontinuities is
considerably slower. Therefore to obtain a more representative displacement, results
were taken from 10 s; all time-displacement plots for data throughout Section 4.3 are
presented for reference, within the following Sections.

Time-displacement plots for purely cohesive discontinuity

Time-Displacement plots for a variety of slope models each


with a different displacement damping (each model has a
cohesion of 35 kPa, and equivilant FOS of 0.96)
0
0 2 4 6 8
-0.002
-0.004
Vertical Displacement (m)

-0.006 damping = 0.068


-0.008 damping = 0.067
-0.01
damping = 0.066
-0.012
damping = 0.065
-0.014
damping = 0.06
-0.016

-0.018
-0.02
Time (s)

Time-Displacement plot for different damped models, all with 35 kPa global cohesion.
The graph illustrates a considerable gap between models with displacement damping
of 0.066 and those with a damping of 0.067, where either a lot or relatively little
displacement occurs respectively.

277
APPENDIX G: Additional results for ELFEN planar failure models

Time-Displacement plots for a variety of slope models each


with a different displacement damping (each model has a
cohesion of 37 kPa, and equivilant FOS of 1.01)
0
0 2 4 6 8
-0.002
-0.004
Vertical Displacement (m)

-0.006 damping = 0.07


-0.008 damping = 0.067

-0.01 damping = 0.066


damping = 0.065
-0.012
damping = 0.064
-0.014
damping = 0.063
-0.016
-0.018
-0.02
Time (s)

Time-Displacement plot for different damped models, all with 39 kPa global cohesion.
This shows that at 37 kPa of global cohesion there is a critical point between 0.064 and
0.063 displacement damping where displacement either occurs or is restrained
respectively. This demonstrates that for an accurate response the damping within this
model should be >0.063, as no vertical displacement should occur if the model has a
FOS of 1.01.

Time-Displacement plots for a variety of slope models each


with a different displacement damping (each model has a
cohesion of 39 kPa, and equivilant FOS of 1.07)
0
0 2 4 6 8
-0.002
-0.004
Vertical Displacement (m)

-0.006 damping = 0.07


-0.008 damping = 0.067

-0.01 damping = 0.066


damping = 0.065
-0.012
damping = 0.064
-0.014
damping = 0.063
-0.016

-0.018
-0.02
Time (s)

Time-Displacement plot for different damped models, all with 39 kPa global cohesion.
This illustrates the anomaly found when displacement damping = 0.064, as a maximum
displacement occurs even through there is a high cohesion and a lower displacement
damping has still restrained from failing (in the case of damping = 0.063).

278
APPENDIX G: Additional results for ELFEN planar failure models

Time-displacement plots for purely frictional discontinuity


The following Figures illustrate the behaviour upon a purely frictional discontinuity; it
can be noted that model successfully reaches equilibrium, only when FOS is <1, with
more rapid complete failure when point damping is low, whereas a slow creep
behaviour with high point damping. In situations where FOS >1, equilibrium has not
been reached within the 10 s runtime.

Time-Displacement Plot for different displacement damped


o
models with 54 friction

0
0 2 4 6 8 10 12

-0.05
Vertical Displacement (m)

-0.1
Damping = 0.125
-0.15 Damping = 0.15
Damping = 0.2
-0.2

-0.25

-0.3
Time (s)

Time-Displacement plot for different damped models, with a global friction of 54o.

Time-Displacement Plot for different displacement damped


o
models with 56 friction

0
0 2 4 6 8 10 12

-0.005
Vertical Displacement (m)

-0.01
Damping = 0.125
-0.015 Damping = 0.15
Damping = 0.2
-0.02

-0.025

-0.03
Time (s)

Time-Displacement plot for different damped models, all with 56o global friction. Here
although the vertical displacement is smaller it can clearly be seen that displacement is
still increasing.

279
APPENDIX G: Additional results for ELFEN planar failure models

Time-Displacement Plot for different displacement damped


o
models with 58 friction

0
0 10 20 30 40 50

-0.0005
Vertical Displacement (m)

-0.001
Damping = 0.125
Damping = 0.15
-0.0015
Damping = 0.2
Damping = 0.5
-0.002

-0.0025

-0.003
Time (s)

Time-Displacement plot for different damped models, all with 58o global friction. Here
although the vertical displacement is smaller it can clearly be seen that displacement is
still increasing. Simulations have in fact been aloud to continue for up to 150 s runtime,
and the displacement has shown to still be linearly increasing.

280
APPENDIX G: Additional results for ELFEN planar failure models

G.2 Data from penalty study (slope 1)

Purely cohesive discontinuity: Purely frictional discontinuity:

Time-Displacement Plot - Time-Displacement Plot -


Pn = 0.3 GPa/m and Pt = 0.03 GPa/m Pn = 0.3 GPa/m and Pt = 0.03 GPa/m

Runtime (s) Runtime (s)


0 5 10
0 5 10
Vertical displacement

Vertical displacement (m)


0
0
-0.5
-0.5 54° (FOS = 0.9)
c = 33 kPa
-1
-1 (FOS = 0.9)
(m)

-1.5 56° (FOS = 1)


-1.5
c = 40 kPa
-2 (FOS = 1.1) -2 58° (FOS = 1.08)

-2.5 -2.5

Time-Displacement Plot - Time-Displacement Plot -


Pn = 3 GPa/m and Pt = 0.3 GPa/m Pn = 3 GPa/m and Pt = 0.3 GPa/m

Runtime (s)
Runtime (s)
0 5 10
0 5 10
0
Vertical displacement (m)

Vertical displacement (m)

0
-0.5
c = 33 kPa -0.5
54 (FOS = 0.9)
-1 (FOS = 0.9) -1
56 (FOS = 1)
-1.5 c = 40 kPa -1.5
(FOS = 1.1) 58 (FOS = 1.1)
-2 -2

-2.5 -2.5

Time-Displacement Plot - Time-Displacement Plot -


Pn = 30 GPa/m and Pt = 3 GPa/m Pn = 30 GPa/m and Pt = 3 GPa/m

Runtime (s)
Runtime (s)
0 5 10
0 10
Vertical displacement

Vertical displacement (m)

0
0 -0.5
-0.5 c = 33 kPa 54° (FOS = 0.9)
-1
-1 (FOS = 0.9)
(m)

56° (FOS = 1)
-1.5 -1.5
c = 40 kPa
58° (FOS = 1.1)
-2 (FOS = 1.1) -2
-2.5 -2.5

281
APPENDIX G: Additional results for ELFEN planar failure models

G.3 Data from penalty study (slope 3)

Purely cohesive discontinuity: Purely frictional discontinuity:

Time-Displacement Plot - Time-Displacement Plot -


Pn = 3 GPa/m and Pt = 0.3 GPa/m Pn = 0.3 GPa/m and Pt = 0.03 GPa/m

Runtime (s)
Runtime (s)
0 5 10
0 5 10

Vertical displacement (m)


Vertical displacement (m)

0
0
-0.5 -0.5
c = 33 kPa
54 (FOS = 0.93)
(FOS = 0.9) -1
-1
c = 36.5 kPa 56 (FOS = 1)
-1.5 (FOS = 1) -1.5
-2 c = 40 kPa -2 58 (FOS = 1.08)
-2.5 (FOS = 1.09)
-2.5

Time-Displacement Plot - Time-Displacement Plot -


Pn = 3 GPa/m and Pt = 0.3 GPa/m Pn = 3 GPa/m and Pt = 0.3 GPa/m

Runtime (s) Runtime (s)


0 5 10 0 5 10
Vertical displacement (m)

Vertical displacement (m)

0 0

-0.5 c = 33 kPa -0.5 54 (FOS = 0.93)


-1 (FOS = 0.9) -1
c = 36.5 kPa 56 (FOS = 1)
-1.5 (FOS = 1) -1.5
-2 c = 40 kPa -2 58 (FOS = 1.08)
(FOS = 1.09)
-2.5 -2.5

Time-Displacement Plot - Time-Displacement Plot -


Pn = 30 GPa/m and Pt = 3 GPa/m Pn = 30 GPa/m and Pt = 3 GPa/m

Runtime (s)
Runtime (s)
0 5 10
0 5 10
Vertical displacement (m)

0
Vertical displacement (m)

0
-0.5
-0.5 c = 33 kPa 54 (FOS = 0.93)
-1 (FOS = 0.9) -1
c = 36.5 kPa 56 (FOS = 1)
-1.5 (FOS = 1) -1.5
-2 c = 40 kPa -2 58 (FOS = 1.08)
(FOS = 1.09)
-2.5
-2.5

282
APPENDIX H: Modification of planar failure model 3

APPENDIX H: Suggested modifications of slope 3


model

Detail and results from modifications suggested to planar failure slope 3 model
presented within Section 4.3.3:

Project file:
1. “slope1h-2_option166,damping,Tcr.elf”

As stated in Section 4.3.3, within the thesis, several modifications were suggested by
Rockfield (2006, personal communication). In addition to the modifications listed in
Section 4.3.3, Rockfield (2006, personal communication) suggested that the slope
model should follow a ‘Constraint release method,’ to minimising dynamic effects. As
outlined in Section 3.1.1, this is not possible in a discontinuity-controlled slope;
subsequently an ‘applied displacement’ approach was used.

However, there was some difficulty with successfully simulating a constraint-release


method using ELFEN pre-processor version 3.8.3. Using this particular code (and
current codes), an applied displacement load had to be used as the constraint, and
with limited guidance this was performed in an incorrect manner resulting in unrealistic
failure of the lines to which it had been applied. Consequently this final modification
could not be included at this stage. (As discussed in Section 4.5 within the thesis, a
constraint-release approach was later implemented following proper guidance and
using a newer version of ELFEN).

The following section presents a penalty study, similar to that reported in Appendix G,
but within a model which included the modifications suggested by Rockfield (2006,
personal communication). The final section presents results from a brief review,
regarding the influence of the ‘fracture opening flag’ and the fracture energy, on net
displacement within a modified model of slope 3.

283
APPENDIX H: Modification of planar failure model 3

Results

Time-displacement plots for purely cohesive discontinuity

Time-Displacement Plot for single-surface two-staged model


with displacement damping, option 166 and a lower T cr ; surface
Pn = 3e7 and Pt = 3e6 Pa/m

Runtime (s)
0
0 5 10
Vertical Displacement (m)

-0.5

c = 33 kPa (FOS = 0.9)


-1
c = 36.5 kPa (FOS = 1)

-1.5

-2

Increasing Penalties (in a single-surface model)


-2.5

Time-Displacement Plot for single-surface two-staged model


with displacement damping, option 166 and a lower T cr ; surface
Pn = 3e8 and Pt = 3e7 Pa/m
Runtime (s)
0
0 5 10

-0.5
Vertical Displacement (m)

c = 33 kPa (FOS = 0.9)


-1 c = 36.5 kPa (FOS = 1)
c = 40 kPa (FOS = 1.09)

-1.5

-2

-2.5

Time-Displacement Plot for single-surface two-staged model


with displacement damping, option 166 and a lower T cr ; surface
Pn = 3e9 and Pt = 3e8 Pa/m
Runtime (s)
0
0 5 10

-0.5
Vertical Displacement (m)

c = 33 kPa (FOS = 0.9)


-1 c = 36.5 kPa (FOS = 1)
c = 40 kPa (FOS = 1.09)

-1.5

-2

-2.5

Figure H.1: The influence of penalties on displacement, upon a purely cohesive discontinuity within
the planar failure slope 3 model that was modified by Rockfield (2006, personal communication).

284
APPENDIX H: Modification of planar failure model 3

Time-displacement plots for purely frictional discontinuity


Time-Displacement Plot for single-surface two-staged model
with displacement damping, option 166 and a lower Tcr ; surface
Pn = 3e7 and Pt = 3e6 Pa/m

Runtime (s)
0
0 5 10
Vertical Displacement (m)

-0.5

Φ = 54 degrees (FOS = 0.93)


-1 Φ = 56 degrees (FOS = 1)
Φ = 58 degrees (FOS = 1.08)
-1.5

-2

Increasing Penalties (in a single-surface model)


-2.5

Time-Displacement Plot for single-surface two-staged model


with displacement damping, option 166 and a lower Tcr ; surface
Pn = 3e8 and Pt = 3e7 Pa/m

Runtime (s)
0
0 5 10

-0.5
Vertical Displacement (m)

Φ = 54 degrees (FOS = 0.93)


-1 Φ = 56 degrees (FOS = 1)
Φ = 58 degrees (FOS = 1.08)

-1.5

-2

-2.5

Time-Displacement Plot for single-surface two-staged model


with displacement damping, option 166 and a lower Tcr ; surface
Pn = 3e9 and Pt = 3e8 Pa/m

Runtime (s)
0
0 5 10
Vertical Displacement (m)

-0.5

Φ = 54 degrees (FOS = 0.93)


-1 Φ = 56 degrees (FOS = 1)
Φ = 58 degrees (FOS = 1.08)
-1.5

-2

-2.5

Time-Displacement Plot for single-surface two-staged model


with displacement damping, option 166 and a lower Tcr ; surface
Pn = 3e10 and Pt = 3e9 Pa/m

Runtime (s)
0
0 5 10
-0.2
Vertical Displacement (m)

-0.4
-0.6 Φ = 54 degrees (FOS = 0.93)
Φ = 56 degrees (FOS = 1)
-0.8
Φ = 58 degrees (FOS = 1.08)
-1

-1.2

-1.4

-1.6

-1.8

Figure H.2: The influence of penalties on displacement, upon a purely frictional discontinuity within
the planar failure slope 3 model, which was modified by Rockfield (2006, personal communication).

285
APPENDIX H: Modification of planar failure model 3

Trials upon a purely frictional slope with a lower Fracture Energy,


with influence of fracture opening flag:

It was noticed during the final calibration models that all previous models had high
fracture energy (see Appendix A). Fracture energy is a parameter that should not
directly influence displacement within a model where there is solely shear displacement
and no fracturing occurring. However, as is illustrated in the results below, net
displacement at 10 s increased dramatically when fracture energy is reduced

In addition the influence of ‘option 166’ (see Appendix A), was determined. In this case
net displacement decreased slightly when option 166 was implemented.

Finally it was noted that there was a slight influence due to deactivating the ‘fracture
opening flag.’ It was suggested by Elmo (2006, personal communication) that a
modification had to be made to the fracture opening flag when using ELFEN pre-
processor version 3.8.3. If the fracture opening flag is active, it causes the
discontinuity to open slightly when the mesh is generated, inserting a red-line to allow
discontinuities to be clearly visible (Rockfield, 2008, personal communication). This
slight opening of discontinuities should have very minimal influence upon the resulting
simulation, however as can be seen by the results presented below, de-activating the
fracture opening flag led to an increase in net displacement at 10 s in this case.

Influence of Fracture Energy, option 166 and fracture opting


flag within slope 3: Time-Displacement Plot for surface Pn = 3e9
o
and Pt = 3e8 Pa/m and Friction = 58

Runtime (s)
0
0 5 10 Fracture Energy = 200 J/m2,
-0.2 no option 166 and fracture
Vertical Displacement (m)

opening flag = 1
-0.4
Fracture Emergy = 20 J/m2; no
-0.6 option 166 and fracture opeing
flag = 1
-0.8
Fracture Emergy = 20 J/m2;
-1 option 166 and fracture opeing
flag = 1
-1.2 Fracture Emergy = 20 J/m2;
option 166 and fracture opeing
-1.4
flag = 0
-1.6

Figure H.3: Influence of fracture energy, option 166 and fracture opening
flag in Planar failure Slope 3 model, with a purely frictional discontinuity.

286
APPENDIX I: Chalk step-path model 1 development

APPENDIX I: Chalk step-path model 1 – detail on model


development and findings

Simulations were conducted on the Chalk step-path model 1, presented in Figure


5.13a, using the mesh control, model parameters and sequence presented within the
model database and the boundaries in Appendix L. However a single excavation
routine was used to unload the slope, as opposed to a constraint release, used in the
other models.

Model results and development


The following section gives detail on efforts to achieve kinematic release and realistic
fracturing within model 1.

Firstly attempts were made to accomplish kinematic release within model 1 using
ELFEN version 3.8.5; however regardless of the discontinuity  applied the basal
fracture, separation at the joint termination could not be achieved, as illustrated in
Figure I.1a. On the addition of ‘Option 166,’ (see Appendix A), a degree of sliding was
permitted on the basal discontinuity. However separation at the joint termination was
still not achieved, resulting in increased deformation at the toe of the slope as
illustrated in Figure I.1b.

Finally sliding on the basal fracture was achieved via extending it into the excavation
block, as illustrated in Figure I.1c, resulting in the kinematic release as illustrated in
Figure 1.1d. This practice could possibly cause problems within larger models, as
discussed in Appendix F. Despite this, further development of the model led to the
implementation of Rayleigh damping within a model simulated using ELFEN
version 3.9.0, as opposed to point damping. This improved fracturing within the model
as presented in Figure I.1e.

287
APPENDIX I: Chalk step-path model 1 development

Figure I.1: Development of ELFEN step-path chalk cliff model 1. (a) and (b) Kinematic release could
not be achieved, regardless of  on basal discontinuity, with ‘option 166’ permitting further
deformation. (c) Separation of basal discontinuity via extension through slope face; consequently
release and improved fractured profile through the implementation of Rayleigh damping (c) and (d)
respectively.

288
APPENDI X J: Data from ELFEN Biaxial modelling

APPENDIX J: Data from ELFEN biaxial modelling


The following appendices presents results from the biaxial modelling in ELFEN, which
was conducted as part of the final case study outlined in Chapter 8.

J.1 Different results from different ELFEN versions


During the primary uniaxial analysis within ELFEN, it was checked whether data
compared well with the uniaxial results for the same model presented within Elmo
(2006). It was discovered that the same results could only be achieved if the same
older version of ELFEN was used, as the new current version of ELFEN returned
significantly different data, as is illustrated in Figure J.1.

Figure J.1: Stress-strain plot illustrating the different results for the same model,
which are obtained when using different versions of ELFEN analysis software.

Following these findings, Rockfield (2008, personal communication) advised that there
have been a lot of changes between these two versions, especially concerning
improvements in the discrete contact and the combined Mohr Coulomb-Rankine
constitutive model. Therefore in order to use the current version of ELFEN, re-
calibration of the model parameters (loading function, damping and normal penalty) is
necessary. Consequently the older analysis version, 3.3.31, was used for the models
throughout Section 8.3, as the model parameters had already been calibrated for this
by Elmo (2006).

289
APPENDI X J: Data from ELFEN Biaxial modelling

J.2 Problem with restart function within ELFEN Analysis


version 3.3.31
When collating the results for ELFEN biaxial simulations, it was discovered that there
was a fundamental problem with using the restart function within Analysis version
3.3.31. Within analysis 3.3.31, a re-initialisation occurs during a restart, causing
dynamics within the model (Rockfield, 2008, personal communication). The
consequence of this is an artificial peak in vertical stress.

Note that it has been advised by Rockfield (2008, personal communication), that this
issue has been addressed within the more current versions of ELFEN and instead
restarting the simulations acts as intended, to allow a temporary pause during the
simulation.

J.3 Biaxial results from non-downgraded models


The biaxial modelling of the Middleton pillar sections, demonstrated that the
Hoptonwood Limestone has a significant mass strength. Figure J.2 presents the data
from the biaxial modelling of section RC; note that the stress-strain plot of the model
with 2 MPa confinement, presented in Figure J.2, appears to not have reached its final
peak. This could also be the case for the 1 MPa confined model.

After the RC simulations, it was noticed the models were set to terminate at 2% strain,
following the conduct of Elmo (2006). The RC simulations should be repeated with a
higher value of maximum axial strain, to ensure the actual σ1max is detected. However
due to the logistical time frame of this research, it was decided that the further
simulations (on RE and the horizontal section) would be set to terminate at a runtime
that enabled a rate of at least 2.5% strain to be observed. Consequently Figures J.3
and J.4 present the data from this testing, with the models run to appropriately higher
strain rates.

290
APPENDI X J: Data from ELFEN Biaxial modelling

Data Set Hoek-Brown mass strength properties


(peaks up gradient (a)
intercept (b) b a
to following
b = sσci2 s= a = mb ci
mb =
axial 2
(MPa2) ci (MPa) ci
strain):
1.5% 37.5 0.0163 349 7.276
2.5% 0 0 484 9.472

 c 1 - sin 
Data Set Mohr-Coulomb mass strength properties
 a-1
1 + sin  = sin -1 
1 - sin 
(peaks up

Gradient (a)
 a + 1
to following Intercept c=

1 - sin
axial a= σc (MPa)
strain): (°) (MPa)
1.5% 11.5 57.1 7.4 1.1
2.5% 13.4 59.3 6.6 0.9

Figure J.2: Stress-strain plot of RC biaxial models under differing confinements, with screen
shots illustrating the state of fracture. Annotation is given to illustrate the fractured state at
peaks in vertical stress that occur up to 1.5% and 2.5% axial strain, which are used to create
corresponding failure envelopes with consequent mass strength derivation (σci = 48 MPa).

291
APPENDI X J: Data from ELFEN Biaxial modelling

Data Set Hoek-Brown mass strength properties


(peaks up gradient (a)
intercept (b) b a
to following
b = sσci2 s= a = mb ci
mb =
axial 2
(MPa2) ci (MPa) ci
strain):
1.5% 63 0.0274 102 2.138
2.5% 224 0.0974 278 5.812

 c 1 - sin 
Data Set Mohr-Coulomb mass strength properties
 a-1
1 + sin  = sin -1 
1 - sin 
(peaks up

Gradient (a)
 a + 1
to following Intercept c=

1 - sin
axial a= σc (MPa)
strain): (°) (MPa)
1.5% 6.2 46.4 6.8 1.4
2.5% 10.3 55.4 11.3 1.8

Figure J.3: Stress-strain plot of RE biaxial models under differing confinements, with screen shots
illustrating the state of fracture. Annotation is given to illustrate the fractured state at peaks in
vertical stress that occur up to 1.5% and 2.5% axial strain, which are used to create corresponding
failure envelopes with consequent mass strength derivation (σci = 48 MPa).

292
APPENDI X J: Data from ELFEN Biaxial modelling

Data Set Hoek-Brown mass strength properties


(peaks up gradient (a)
intercept (b) b a
to following
b = sσci2 s= a = mb ci
mb =
axial 2
(MPa2) ci (MPa) ci
strain):
1.5% 162 0.0704 303 6.881
2.5% 0 0.0000 867 18.064*

 c 1 - sin 
Data Set Mohr-Coulomb mass strength properties
 a-1
1 + sin  = sin -1 
1 - sin 
(peaks up

Gradient (a)
 a + 1
to following Intercept c=

1 - sin
axial a= σc (MPa)
strain): (°) (MPa)
1.5% 11.8 55.2 10.2 1.8
2.5% 9.6 63.4 9.6 1.1

Figure J.4: Stress-strain plot of horizontal biaxial models under differing confinements, with
screen shots illustrating the state of fracture. Annotation is given to illustrate the fractured state at
peaks up to 1.5 % and 2.5 % axial strain, which are used to create corresponding failure
envelopes and consequent mass strength derivation (σci = 48 MPa).
* mb value very high, due to high strength at high confinement (result not included in main thesis).

293
APPENDI X J: Data from ELFEN Biaxial modelling

J.4 Biaxial results from downgraded models


Following degradation of intact strength properties, RE and horizontal sections were re-
analysed to give the data presented in J.5 and J.6.

Data Set Hoek-Brown mass strength properties


(peaks up gradient (a)
intercept (b) b a
to following
b = sσci2 s= a = mb ci
mb =
axial 2
(MPa2) ci (MPa) ci
strain):
1.5% 0.0 0.0000 125 2.553
2.5% 0.9 0.0004 201 4.180

 c 1 - sin 
Data Set Mohr-Coulomb mass strength properties
 a-1
1 + sin  = sin -1 
1 - sin 
(peaks up

Gradient (a)
 a + 1
to following Intercept c=

1 - sin
axial a= σc (MPa)
strain): (°) (MPa)
1.5% 4.1 48.4 6.9 0.771
2.5% 5.2 52.5 8.7 0.876

Figure J.5: Stress-strain plot of downgraded (quartered intact strength) RE biaxial models under
differing confinements, with screen shots illustrating the state of fracture. Annotation is given to
illustrate the fractured state at peaks up to 1.5 % and 2.5 % axial strain, which are used to create
corresponding failure envelopes (σci = 12 MPa).

294
APPENDI X J: Data from ELFEN Biaxial modelling

Data Set Hoek-Brown mass strength properties


(peaks up gradient (a)
intercept (b) b a
to following
b = sσci2 s= a = mb ci
mb =
axial 2
(MPa2) ci (MPa) ci
strain):
1.5% 0.0 0.0000 575 11.969
2.5% 956 0.0130 960 19.971*

 c 1 - sin 
Data Set Mohr-Coulomb mass strength properties
 a-1
1 + sin  = sin -1 
1 - sin 
(peaks up

Gradient (a)
 a + 1
to following Intercept c=

1 - sin
axial a= σc (MPa)
strain): (°) (MPa)
1.5% 15.5 61.5 6.4 0.813
2.5% 19.5 64.4 9.5 1.083
Figure J.6: Stress-strain plot of downgraded (quartered intact strength) horizontal biaxial models
under differing confinements, with screen shots illustrating the state of fracture. Annotation is given
to illustrate the fractured state at peaks up to 1.5 % and 2.5 % axial strain, which are used to create
corresponding Hoek-Brown failure envelopes (σci = 12 MPa).
* mb value very high, due to high strength at high confinement (result not included in main thesis).

295
APPENDI X J: Data from ELFEN Biaxial modelling

Depending on the results that were used, different envelopes and subsequent mass
strengths can be derived for the equivalent continuum. This was reviewed in the
downgraded model, but only using the Mohr-Coulomb criterion due to time constraints.
The results for this are presented in Figure J.7 below.

Mohr-Coulomb mass strength properties for downgraded

 c 1 - sin 
dominantly vertically fractured (RE) mass
-1  a - 1 
Data taken from

1 + sin  = sin  a + 1  Intercept σc c = 1 - sin 


following range of σ3 to Gradient (a)

1 - sin
create envelope (MPa):
a= (MPa)
(°) (MPa)
0, 0.5, 1, 2 and 4. 12.0 58 2.8 0.410
0, 1, 2 and 4. 12.5 58 1.2 0.164
0, 0.5, 1 and 2. 8.7 53 5.2 0.876
0, 1 and 2. 9.4 54 3.5 0.577

Mohr-Coulomb mass strength properties for downgraded

 c 1 - sin 
dominantly horizontally fractured mass
-1  a - 1 
Data taken from

1 + sin  = sin  a + 1  Intercept σc c = 1 - sin 


following range of σ3 to Gradient (a)

1 - sin
create envelope (MPa):
a= (MPa)
(°) (MPa)
0, 0.5, 1 and 2. 19.5 64 9.5 1.083
0, 1 and 2. 21.4 66 5.0 0.546

Figure J.7: Mohr-Coulomb failure envelopes and consequent mass strength parameters, derived from a
range of confinements with the original biaxial ELFEN simulations

296
APPENDIX K: Groundwater properties in ELFEN models

APPENDIX K - Derivation of parameters in groundwater


models
In all models with groundwater, the groundwater option within the neutral file (see
Appendix A), was set as 1 (“drained analysis with non-zero pore pressures, based on
total stress”).

Also the basic properties from all models were taken from example of a saturated chalk
slope, given by Rockfield (2007, personal communication).

When the actual values for each parameter are derived, it is clear that there are some
inconsistencies (as presented in Table K.1). In particular the bulk modulus of the fluid
(Kf) may have been lowered by Rockfield (2007, personal communication) in the
example slope, to make the simulation quicker. Ideally given time, all of the models
need re-running with the values (presented in Table K.1), which can be derived using
the below equations.

True density
When modelling groundwater within ELFEN, the true density of the solid () has to be

 = (1 -  )  s +  f
modified through the following calculation:
[K.1]

Subsequently the solid (in the groundwater specification within the neutral file) should
be entered as the previous true density.

Bulk modulus of the solid


Bulk modulus of the solid (Ks) can be derived from the following equation and calculus:

3(1 - 2 )
E
Ks = [K.2]

Table K.1: Inconsistencies noted after simulations when model parameters were checked. The
parameters used in the models were from an example of a saturated chalk slope, given by Rockfield
(2007, personal communication); however when model parameters are derived using above formulae,
it is clear that values should have been different.

Parameter Plane failure Chalk Delabole


Value used in
2500 1490 2500
True density simulation
(kg/m3) Value that should have
2200 1700 2845
been used
Value used in
Density of 2200 1700 2845
simulation
the solid
3 Value that should have
(kg/m ) 2500 1490 2500
been used
Bulk Value used in
0.2 0.2 0.2
modulus of simulation
the fluid Value that should have
2 2 2
(GPa) been used
Bulk Value used in
31 31 31
modulus of simulation
the solid Value that should have
13 0.64 13
(GPa) been used

297
APPENDIX L: Additional detail on ELFEN models

APPENDIX L – Model boundaries and mesh details for ELFEN models

Mesh Sphere Detail and project file(s):


Model Model Dimensions Interior density Radius for interior Exterior density Radius for exterior
Location (x, y)
(m2) circle (m) (m2) circle (m)
no mesh spheres
Damping study (Model 1):
“slope1b.elf” – used for purely cohesive discontinuity point damping study
“slope1a-1.elf” to “slope1a-3.elf” – used for purely frictional discontinuity point damping and penalty
Planar failure models
studies
1 and 2
Penalty study (Model 2):
“slope1g.elf” – for simulations where there is a purely frictional discontinuity
“slope1g-1.elf” – for simulations where there is a purely cohesive discontinuity

no mesh spheres

Planar failure model 3 “slope1g.elf” – for simulations where there is a purely frictional discontinuity
“slope1g-1.elf” – for simulations where there is a purely cohesive discontinuity

no mesh spheres

“drop_load_constraint_release.elf” – for purely frictional simulations


Planar failure model 4
“2-1drop_load_constraint_release_cohesional.elf” for purely cohesive and mixed cohesive-frictional
simulations
298
APPENDIX L: Additional detail on ELFEN models

Mesh Sphere Detail and project file(s):


Model Model Dimensions Interior density Radius for interior Exterior density Radius for exterior
Location (x, y)
(m2) circle (m) (m2) circle (m)
no mesh spheres

Planar failure model 5 “drop_load_constraint_release.elf”

7.8, 7.6 Varied between Varied between 0.3 4.2


0
11.1, 0.0 0.3 and 0.6 and 0.6 4.2
Hutchinson Joss Bay model
1
“2tension_crack(excavation).elf”

7.8, 7.6 0.2 0.4 4.2


0
11.1, 0.0 0.2 0.4 4.2
Hutchinson Joss Bay model
“new_hutchinson_tension_crack_stress_monitoring.elf” and
2
“new_hutchinson_low_TensileStrength.elf”

7.8, 7.6 0.2 0.4 4.2


11.1, 0.0 0.2 0 0.4 4.2
Hutchinson Joss Bay model 6.0, 14.0 0.2 0.4 4.2
3
“final models” simulation set –
“1no_tension_crack(excavation).elf”
7.8, 7.6 0.2 0.4 4.2
11.1, 0.0 0.2 0 0.4 4.2
Hutchinson Joss Bay model 6.0, 14.0 0.2 0.4 4.2
4
“final models” simulation set –
“4no_tension_crack(no_excavation)-with_option37.elf”
299
APPENDIX L: Additional detail on ELFEN models

Mesh Sphere Detail and project file(s):


Exterior density Radius for
Model Model Dimensions Location (x, Interior Radius for
2 (m2) exterior circle
y) density (m ) interior circle (m)
(m)
7.8, 7.6 0.2 0.4 4.2
11.1, 0.0 0.2 0 0.4 4.2
6.0, 14.0 0.2 0.4 4.2
Hutchinson Joss Bay
model 5
“final models” simulation set –
“5-2lower_dem_friction.elf”

7.8, 7.6 0.2 0.4 4.2


11.1, 0.0 0.2 0 0.4 4.2
6.0, 14.0 0.2 0.4 4.2
Hutchinson Joss Bay
model 6
“final models” simulation set –
“6vertical_excavation4stages.elf”

7.8, 7.6 0.2 0.4 4.2


11.1, 0.0 0.2 0 0.4 4.2
6.0, 14.0 0.2 0.4 4.2
Hutchinson Joss Bay
model 7
“final models” simulation set –
“7horizontal_excavation4stages - geometry and mesh.elf”

no mesh spheres

Model 1: “2dry_fractured_chalk.elf” and “3, 3-1 and 4dry_fractured_chalk.elf”


Chalk Step-path models Model 2: “6-1new_geom_chalk_constraint_release.elf”
Model 3: “7-1steeper_slope_chalk_step_constraint_release.elf”
Model 4: “8-1shallower_stepped_chalk_constraint_release.elf”
300
APPENDIX L: Additional detail on ELFEN models

Mesh Sphere Detail and project file(s):


Model Model Dimensions Interior density Radius for interior Exterior density Radius for exterior
Location (x, y)
(m2) circle (m) (m2) circle (m)
detail of mesh sphere is unknown, as mesh file was lost during file transfer

Delabole model 1 “10-1delabole_excavation_mohr-rankine - geometry and mesh.elf” (neu number 3)

75, 85 2 0 6 10

“Mohr-Rankine large model” simulation set –


Delabole model 2
“1delabole_with_fractures - profile and mesh.elf” (neu number 1)

detail of mesh sphere is unknown, as mesh file was lost during file transfer

“extended claylodes” simulation set -


Delabole model 3
“5delabole_extended(no_excavation).elf” (neu number 1)

86, 77 1 0 2 2
82, 77.5 1 0 2 2
65, 121 1 0 2 2
69.5, 103.5 1 0 2 3
Delabole models 4 to 6
76.5, 88 1 0 2 4
Model 4: “1-1delabole_rankine” (neu number 10)
Model 5*: “7-1delabole_rankine - geometry and mesh” (neu number 2)
Model 6: “6-1delabole_rankine.elf”

* Delabole model 5 had a coarser mesh density within mesh spheres (unfortunately data was lost in file transfer).
301
APPENDIX L: Additional detail on ELFEN models

Mesh Sphere Detail and project file(s):


Model Model Dimensions Interior density Radius for interior Exterior density Radius for exterior
Location (x, y)
(m2) circle (m) (m2) circle (m)
no mesh spheres

“extended claylodes” simulation set -


Delabole model 7
“5delabole_extended(no_excavation).elf” (neu number 3)

no mesh spheres
Model 8: “extended claylodes” simulation set -
“6delabole_extended_shortahs - profile and mesh.elf”
Delabole model 8 and 9
Model 9: from “extended claylodes” simulation set –
“8delabole_ extended_shortahs.elf” – model with rotation of upper active wedge
“10delabole_extended_shortahs.elf” – model with rotation of lower active wedge
86, 77 1.5 0 2 3
82, 75.5 1.5 0 2 3
72, 94 1.5 0 2 3
Delabole model 10
“extended claylodes” simulation set -
“9delabole_ extended_shortahs.elf”
no mesh spheres

Delabole model 11 “planar failure” simulation set -


“5-1longer_constraint_release.elf” (neu number 6)

386, 327 0 2 1.5 3


382, 325.5 0 2 1.5 3
Delabole model 12 “step-path wet models” simulation set –
“4-1simple_step-path_model_crelease.elf” (neu number 5)
302
APPENDIX L: Additional detail on ELFEN models

Mesh Sphere Detail and project file(s):


Radius for Exterior Radius for
Model Model Dimensions Interior
Location (x, y) 2 interior circle density (m2) exterior circle
density (m )
(m) (m)
386, 327 0 2 1.5 3
382, 325.5 0 2 1.5 3
Delabole model
13* from “step-path” simulation set –
“5-1step-path_model_crelease_extended_claylodes.elf” (neu number 2)

*Additional detail for Delabole model 13:


Simulations were run using same analysis version, options, material parameters and control as those presented for model 12. However it was found
that high  45° and high penalties (Pn = 30GPa/m, Pt = 10GPa/m), were required on the claylodes throughout the whole runtime of the model to
prevent large-scale failure between the clyalodes:

Reduction of strength upon the shortah and base-plane was still included in a post-failure stage, to allow
unhindered sliding on discontinuities and consequent comminution and deposition (the Pn and Pt were still
reduced to 20MPa/m and 2MPa/m respectively, and the  to 36°).

Figure L.1: Extensive failure when weak properties are applied to claylodes.
303
SECTION OF DATABASE: i
Particular model (referred to in thesis): Planar failure slope 1 Planar failure slope 2 Planar failure slope 3 Planar failure slope 4
Altering the damping to Effect of damping within a Effect of penalties within a Testing the groundwater
Model Intent
attain limit states larger-scale model single-surface model module within ELFEN

ELFEN Analysis Version 3.3.40 3.3.40 3.3.40 3.9.1


Additional 'option(s)' - see Appendix A - - - 166 and 37
K-ratio not applied not applied not applied 0.33
Surface Domain 
Main Failure Main Failure Single surface model Single surface model

Control  Parameter 
Slope Block Slope Block (embedded discontinuity) (embedded discontinuity)
Value Value Value Value Value Value
Mass-strength and related properties

Young’s Modulus (GPa ) 50 20 50 20 20 20


Main
Poisson's ratio 0.2 0.2 0.2 0.2 0.25 0.25
properties
Density (kg/m3) 2600 2600 2600 2600 2500 2500
Cohesion (MPa) 9
Mohr-
Friction Angle (°) Purely Rankine Purely Rankine Purely Rankine 54
Coulomb and
Dilation angle (°) 5
Rotating
Tensile strength (MPa) 20 20 20 20 5 5
crack 2
Fracture energy (J/m ) 200 200 200 200 200 20
Porosity (%) 20
Density of fluid (kg/m 3) 1000
3
Groundwater Density of solid (kg/m ) Dry model Dry model Dry model 2200
Bulk modulus of fluid (GPa) 0.2
Bulk modulus of solid (GPa) 30
Viscosity (Pa s) 0.001
Contact damping (%) 30 30 30 10
Default discontinuity and mesh-

Field (m) 0.1 0.1 0.1 0.1


Normal penalty (GPa/m) 30 10 Varied: 0.3, 3 and 30 0.2
Discrete Tangential penalty (GPa/m) 3 1 Varied: 0.03, 0.3 and 3 0.02
based properties

global Zone (m) 1 1 1 0.29


properties Smallest element (m) 1 1 1 0.5
Friction (°) Varied: 54, 56, 58 and 60 0 Varied: 54, 56, 58 and 60 Varied: 54, 56 and 58
Varied: 0.01, 0.02, 0.03, Varied: 0.01, 0.02, 0.03, Varied: 0.033, 0.0365 and
Cohesion (MPa) 105
0.04 and 0.05 0.04 and 0.05 0.04
Particular discontinuity - - - -
Particular Friction Angle (°) - - - -
discontinuity Cohesion (MPa) - - - -
strength Normal penalty (GPa/m) - - - -
- - - -
Number of stages 
Tangential penalty (GPa/m)

Stage Number  Stage Number  Stage Number  Stage Number 


1* 1 2 3*

1 1 1 2 1 2
Model Sequence

Point damping (%) 10 Varied – see Figure 3.4 10 0 5 5


Critical Time step 0.8 0.8 0.8 0.8 0.9 0.25
Release of
Global gravity Global gravity
Process Global gravity loading Global gravity loading Failure slope
loading loading
constraint
Load/release curve Drop load Drop load Drop load - Drop load Sine-curve
Duration of stage (s) - - - 0.5 s to 9.5 - 0 to 1
Release on ramped-curve
Discontinuity release detail not applied not applied not applied
from: 0 to 0.5 s
Runtime of simulation on 2.8 GHz processor with
2 7* 2 5
512 Mb RAM (min)
Mesh density at boundaries 2 6 1 5
Mesh density applied to
1 3 1 0.5
failure block
Mesh spheres 0 0 0 0
Meshing Algorithm Delaunay Delaunay Delaunay Advancing Front
Geometry and *
mesh detail
10 m

30 m

10 m

20 m

Mesh screen-shot
illustrating model extent

Type of model Failure due to gravity Failure due to gravity Failure due to gravity Release of slope constraint

Degree of point damping Exponential relationship Realistic when on cohesive Successful when penalties
is highly influential upon a found between degree of discontinuity are 20 MPa/m and
purely cohesive surface; point damping and Pn = 3 GPa/m; whereas the 2 MPa/m (Pn and Pt
Result
for a purely frictional displacement on a planar on a purely frictional respectively), with failure
surface effect of damping surface (within a large- discontinuity Pn should be when FOS in
was less critical scale model) 0.3 GPa/m RocPlane is < 0.8

* This was for the fine * Stage 3 has the rise of


* A further stage was mesh; for a coarse mesh groundwater, ramped from
added in penalty study, simulation was less than 1
Additional comments 1.5 to 2 s (with same point
following same sequence minute and mesh
as in slope 3 screenshot was as in damping and time step as
slope 1 model stage 2)
304
SECTION OF DATABASE: i ii
Particular model (referred to in thesis): Planar failure slope 5 Hutchinson Joss Bay model 1 Hutchinson Joss Bay model 2
Model with tension crack to test Model with showing progressive Model with tension crack used for stress
Model Intent
groundwater ELFEN module development of shear strain path monitoring
ELFEN Analysis Version 3.9.1 3.8.5 3.8.5
Additional 'option(s)' - see Appendix A 166 and 37 37 37
K-ratio 0.33 0.32 0.32
Surface Domain 
Single surface model Single surface model Single surface model

Control  Parameter 
(embedded discontinuity) (embedded discontinuity) (embedded discontinuity)
Value Value Value
Mass-strength and related properties

Young’s Modulus (GPa ) 20 1 1


Main
Poisson's ratio 0.25 0.24 0.24
properties
Density (kg/m3) 2500 1700 1700
Cohesion (MPa) 9 0.026 0.026
Mohr-
Friction Angle (°) 54 50 50
Coulomb and
Dilation angle (°) 5 5 5
Rotating
Tensile strength (MPa) 5 0.013 0.013
crack
Fracture energy (J/m2) 20 47.5 47.5
Porosity (%) 20
Density of fluid (kg/m 3) 1000
3
Groundwater Density of solid (kg/m ) 2200 Dry model Dry model
Bulk modulus of fluid (GPa) 0.2
Bulk modulus of solid (GPa) 30
Viscosity (Pa s) 0.001
Contact damping (%) 10 30 30
Default discontinuity and mesh-

Field (m) 0.1 0.04 0.04


Normal penalty (GPa/m) 0.2 0.1 0.1
Discrete Tangential penalty (GPa/m) 0.02 0.01 0.01
based properties

global Zone (m) 0.29 6 6


properties Smallest element (m) 0.5 0.2 0.2
Friction (°) Varied: 54, 56 and 58 30 30
Cohesion (MPa) Varied: 0.033, 0.0365 and 0.04 0 0
Particular discontinuity - Tension crack  Tension crack 
Particular Friction Angle (°) - 60 60
discontinuity Cohesion (MPa) - 0.026 0.026
strength Normal penalty (GPa/m) - 0.1 0.1
- 0.01 0.01
Number of stages 
Tangential penalty (GPa/m)

Stage Number  Stage Number  Stage Number 


3* 2 5*

1 2 1 2 1 2
Model Sequence

Point damping (%) 5 5 5 5 5 5


Critical Time step 0.9 0.25 0.75 0.25 0.75 0.25
Removal of
Global gravity Removal of Global gravity Global gravity Removal of
Process excavation
loading excavation object loading loading excavation object
object
Load/release curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve
Duration of stage (s) 0 to 1 2 to 3 0 to 1 2 to 5 0 to 1 2 to 4
Release on ramped-curve from: Release on ramped-curve from: Release on ramped-curve from:
Discontinuity release detail 1 to 1.2 s 1 to 2 s 1 to 2 s
Runtime of simulation on 2.8 GHz processor with
7 15 80
512 Mb RAM (min)
Mesh density at boundaries 5 10 3
Mesh density applied to
0.5 0.5 3
discontinuities
Mesh spheres 0 2 2
Meshing Algorithm Advancing Front Advancing Front Advancing Front
Geometry and
mesh detail
20 m

45 m

45 m

Mesh screen-shot
illustrating model extent

Type of model Single excavation object Single excavation object Four staged excavation

Successful (with failure between


Model was eventually successful,
0.8 and 0.9 FOS from Model showing dependency on tensile
after a fine enough mesh density
Plane_failure) once phreatic strength, with rigid behaviour when high
Result was found to allow full
surface was modified from initial and more ductile realistic behaviour (for
development of shear strain, and
simplified surface to match chalk) when tensile strength is low
option 37 to allow deformation
triangular pressure distribution

* Stage 3 has the rise of


* Stages 2 to 5 are the same, each with
groundwater, ramped from 3 to 3.5
Additional comments a 2 s duration during which constraint
s (with same point damping and
from each excavation object is relaxed
time step as stage 2)
305
SECTION OF DATABASE: ii
Particular model (referred to in thesis): Hutchinson Joss Bay model 3 Hutchinson Joss Bay model 4 Hutchinson Joss Bay model 5
Simulation of shear strain and Test to analyse the influence of Test to analyse the influence of
Model Intent
subsequent tension crack failure due to gravity on stress state single excavation on stress state

ELFEN Analysis Version 3.8.5 3.8.5 3.8.5


Additional 'option(s)' - see Appendix A 37 37 37
K-ratio 0.32 0.32 0.32
Surface Domain 
Single surface model Single surface model Single surface model

Control  Parameter 
(no discontinuity) (no discontinuity) (no discontinuity)
Value Value Value
Mass-strength and related properties

Young’s Modulus (GPa ) 1 1 1


Main
Poisson's ratio 0.24 0.24 0.24
properties
Density (kg/m3) 1700 1700 1700
Cohesion (MPa) 0.026 0.026 0.026
Mohr-
Friction Angle (°) 50 50 50
Coulomb and
Dilation angle (°) 5 5 5
Rotating
Tensile strength (MPa) 0.013 0.013 0.013
crack
Fracture energy (J/m2) 47.5 47.5 47.5
Porosity (%)
Density of fluid (kg/m 3)
3
Groundwater Density of solid (kg/m ) Dry model Dry model Dry model
Bulk modulus of fluid (GPa)
Bulk modulus of solid (GPa)
Viscosity (Pa s)
Contact damping (%) 25 25 25
Default discontinuity and mesh-

Field (m) 0.04 0.04 0.04


Normal penalty (GPa/m) 0.1 0.1 0.1
Discrete Tangential penalty (GPa/m) 0.01 0.01 0.01
based properties

global Zone (m) 6 6 6


properties Smallest element (m) 0.2 0.2 0.2
Friction (°) 60 25 25
Cohesion (MPa) 0.0264 0 0
Particular discontinuity  - - -
Particular Friction Angle (°) - - -
discontinuity Cohesion (MPa) - - -
strength Normal penalty (GPa/m) - - -
- - -
Number of stages 
Tangential penalty (GPa/m)

Stage Number  Stage Number  Stage Number 


2 2 2

1 2 1 2 1 2
Model Sequence

Point damping (%) 5 5 5 5 5 5


Critical Time step 0.75 0.25 0.75 0.25 0.75 0.25
Removal of Removal of
Global gravity Global gravity Release of Global gravity
Process excavation excavation
loading loading discontinuities loading
object object
Load/release curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve
Duration of stage (s) 0 to 1 2 to 5 0 to 1 1 to 2 0 to 1 2 to 4
Release on ramped-curve from: Release on ramped-curve from: Release on ramped-curve from:
Discontinuity release detail 1 to 2 s 1 to 2 s 1 to 2 s
Runtime of simulation on 2.8 GHz processor with
90 37 65
512 Mb RAM (min)
Mesh density at boundaries 10 10 10
Mesh density applied to
- - -
discontinuities
Mesh spheres 3 3 3
Meshing Algorithm Advancing Front Advancing Front Advancing Front
Geometry and
mesh detail
45 m
75 m

75 m

75 m

Mesh screen-shot illustrating


model extent

Type of model Single excavation object Failure due to gravity Single excavation object

Successful development of tension


crack as a result of progressive Successful with failure noted at both No failure at stress monitoring
Result
development of shear strain from monitoring points A and B point B (within the top of the slope)
notch at slope base

Additional comments
306
SECTION OF DATABASE: ii iii
Particular model (referred to in thesis): Hutchinson Joss Bay model 6 Hutchinson Joss Bay model 7 Chalk Step path models
Test to analyse the influence of
Test to analyse the influence of Development of 4 models, to simulate
Model Intent horizontal excavation on stress
vertical excavation on stress state step-path failure due to groundwater
state
ELFEN Analysis Version 3.8.5 3.8.5 3.9.0
Additional 'option(s)' - see Appendix A 37 37 166 and 37
K-ratio 0.32 0.32 0.32
Surface Domain 
Single surface model Single surface model Single surface model

Control  Parameter 
(no discontinuity) (no discontinuity) (embedded discontinuities)
Value Value Value
Mass-strength and related properties

Young’s Modulus (GPa ) 1 1 1


Main
Poisson's ratio 0.24 0.24 0.24
properties
Density (kg/m3) 1700 1700 1700
Cohesion (MPa) 0.026 0.026 0.026
Mohr-
Friction Angle (°) 50 50 50
Coulomb and
Dilation angle (°) 5 5 5
Rotating
Tensile strength (MPa) 0.013 0.013 0.013
crack
Fracture energy (J/m2) 47.5 47.5 47.5
Porosity (%) 30
Density of fluid (kg/m 3) 1000
3
Groundwater Density of solid (kg/m ) Dry model Dry model 1490
Bulk modulus of fluid (GPa) 0.2
Bulk modulus of solid (GPa) 30
Viscosity (Pa s) 0.001
Contact damping (%) 25 25 25
Default discontinuity and mesh-

Field (m) 0.04 0.04 0.04


Normal penalty (GPa/m) 0.1 0.1 0.02
Discrete Tangential penalty (GPa/m) 0.01 0.01 0.02
based properties

global Zone (m) 6 6 6


properties Smallest element (m) 0.2 0.2 0.5
Friction (°) 25 25 20
Cohesion (MPa) 0 0 0
Particular discontinuity  - - -
Particular Friction Angle (°) - - -
discontinuity Cohesion (MPa) - - -
strength Normal penalty (GPa/m) - - -
- - -
Number of stages 
Tangential penalty (GPa/m)

Stage Number  Stage Number  Stage Number 


5* 5* 3

1 2 1 2 1 2
Model Sequence

Point damping (%) 5 5 5 5 5 5


Critical Time step 0.75 0.25 0.75 0.25 0.75 0.25
Removal of
Global gravity Global gravity Removal of Global gravity Release of slope
Process excavation
loading loading excavation object loading constraint
object
Load/release curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve Drop load Sine-curve
Duration of stage (s) 0 to 1 2 to 4 0 to 1 2 to 4 - 1 to 3
Release on ramped-curve from: Release on ramped-curve from: Release on ramped-curve from:
Discontinuity release detail 1 to 2 s 1 to 2 s 0.5 to 1 s
Runtime of simulation on 2.8 GHz processor with
110 84 10
512 Mb RAM (min)
Mesh density at boundaries 10 10 10
Mesh density applied to
- - 5
discontinuities
Mesh spheres 3 3 0
Meshing Algorithm Advancing Front Advancing Front Advancing Front
Geometry and
mesh detail
75 m

75 m

45 m

Mesh screen-shot illustrating


model extent

Type of model Four staged (vertical) excavation Four staged (horizontal) excavation

Models 1 to 3 were unsuccessful (with


No failure at stress monitoring
either failure prior to rise in
point A (within the toe of the slope) No failure at stress monitoring point
Result groundwater, or no failure), finally
or point B (within the top of the A (within the toe of the slope)
achieved failure due to groundwater
slope)
within model 4

* Stages 2 to 5 are the same, each * Stages 2 to 5 are the same, each * Stage 3 has the rise of groundwater,
with a 2 s duration during which with a 2 s duration during which ramped from 4 to 5.5 s (with same
Additional comments
constraint from each excavation constraint from each excavation point damping and time step as in
object is relaxed object is relaxed stage 2)
307
SECTION OF DATABASE: iv
Particular model (referred to in thesis): Delabole model 1 Delabole model 2 Delabole model 3
Simulation of discontinuity Simulation of discontinuity Simulation of discontinuity
Model Intent
extension across rock-bridge extension across rock-bridge extension across rock-bridge
ELFEN Analysis Version 3.8.5 3.8.5 3.8.5
Additional 'option(s)' - see Appendix A 37 37 37
K-ratio 0.33 0.33 0.33
Surface Domain 
Single surface model Single surface model Single surface model

Control  Parameter 
(embedded discontinuities) (embedded discontinuities) (embedded discontinuities)
Value Value Value
Mass-strength and related properties

Young’s Modulus (GPa ) 20 20 20


Main
Poisson's ratio 0.25 0.25 0.25
properties
Density (kg/m3) 2500 2500 2500
Cohesion (MPa) 2 2 2
Mohr-
Friction Angle (°) 46 46 46
Coulomb and
Dilation angle (°) 5 5 5
Rotating
Tensile strength (MPa) 0.9 0.756 0.9*
crack
Fracture energy (J/m2) 20 20 20
Porosity (%)
Density of fluid (kg/m 3)
3
Groundwater Density of solid (kg/m ) Dry model Dry model Dry model
Bulk modulus of fluid (GPa)
Bulk modulus of solid (GPa)
Viscosity (Pa s)
Contact damping (%) 30 10 10
Default discontinuity and mesh-

Field (m) 0.4 0.4 1


Normal penalty (GPa/m) 1 1 0.1
Discrete Tangential penalty (GPa/m) 0.1 0.1 0.1
based properties

global Zone (m) 7 40 7


properties Smallest element (m) 2 2 10
Friction (°) 31 31 31
Cohesion (MPa) 0.01 0.01 0
Particular discontinuity  - - -
Particular Friction Angle (°) - - -
discontinuity Cohesion (MPa) - - -
strength Normal penalty (GPa/m) - - -
- - -
Number of stages 
Tangential penalty (GPa/m)

Stage Number  Stage Number  Stage Number 


2 2 1

1 2 1 2 1
Model Sequence

Point damping (%) 5 5 5 5 30


Critical Time step 0.75 0.25 0.75 0.25 0.75
Removal of Removal of
Global gravity Global gravity
Process excavation excavation Global gravity loading
loading loading
object object
Load/release curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve Ramp-curve
Duration of stage (s) 0 to 1 2 to 4 0 to 1 2 to 4 0 to 1
Release on ramped-curve from: Release on ramped-curve from: Release on ramped-curve from:
Discontinuity release detail 1 to 2 s 1 to 2 s 4 to 5 s
Runtime of simulation on 2.8 GHz processor with
Between 2 and 5 5 360
512 Mb RAM (min)
Mesh density at boundaries 10 100 30
Mesh density applied to
5 10 1, 5 and 10*
discontinuities
Mesh spheres 1 1 ? geometry file not present ?
Meshing Algorithm Advancing Front Advancing Front Delaunay
Geometry and
mesh detail
150 m

600 m

400 m

Mesh screen-shot illustrating


model extent

Type of model Single excavation object Single excavation object Failure due to gravity

Kinetically unstable model with More kinetically stable, with


Promotion of instability by
fracture prior to release of extended boundaries, also almost
extension of claylodes, resulting in
Result excavation object and finally complete fracture through rock-
successful rock-bridge failure but
large-scale fracture following bridge however then model
no kinematic release
failure becomes stable

*1 m mesh density applied to


stepped base plane, 5 m mesh
Additional comments
density applied to shortah and
10 m applied to claylodes
308
SECTION OF DATABASE: iv
Particular model (referred to in thesis): Delabole model 4 Delabole model 5 Delabole model 6
Simulation of failure mechanism Simulation of failure mechanism Simulation of failure mechanism
Model Intent
along step-path along step-path along step-path

ELFEN Analysis Version 3.8.5 3.8.5 3.8.5


Additional 'option(s)' - see Appendix A 37 37 37
K-ratio 0.33 0.33 0.33
Surface Domain 
Single surface model Single surface model Single surface model

Control  Parameter 
(embedded discontinuities) (embedded discontinuities) (embedded discontinuities)
Value Value Value
Mass-strength and related properties

Young’s Modulus (GPa ) 20 20 20


Main
Poisson's ratio 0.25 0.25 0.25
properties
Density (kg/m3) 2500 2500 2500
Cohesion (MPa)
Mohr-
Friction Angle (°) Purely Rankine Purely Rankine Purely Rankine
Coulomb and
Dilation angle (°)
Rotating
Tensile strength (MPa) 2 2 2
crack
Fracture energy (J/m2) 20 20 20
Porosity (%)
Density of fluid (kg/m 3)
3
Groundwater Density of solid (kg/m ) Dry model Dry model Dry model
Bulk modulus of fluid (GPa)
Bulk modulus of solid (GPa)
Viscosity (Pa s)
Contact damping (%) 10 10 10
Default discontinuity and mesh-

Field (m) 0.2 0.2 0.2


Normal penalty (GPa/m) 1 0.1 0.1
Discrete Tangential penalty (GPa/m) 0.1 0.01 0.1
based properties

global Zone (m) 7 4 7


properties Smallest element (m) 10 5 10
Friction (°) 31 31 31
Cohesion (MPa) 0 0 0
Particular discontinuity  - - -
Particular Friction Angle (°) - - -
discontinuity Cohesion (MPa) - - -
strength Normal penalty (GPa/m) - - -
- - -
Number of stages 
Tangential penalty (GPa/m)

Stage Number  Stage Number  Stage Number 


2 2 2

1 2 1 2 1 2
Model Sequence

Point damping (%) 5 5 5 5 5 5


Critical Time step 0.25 0.18 0.25 0.25 0.75 0.25
Global gravity Global gravity Global gravity
Process Failure Failure Failure
loading loading loading
Load/release curve Ramp-curve - Ramp-curve - Ramp-curve -
Duration of stage (s) 0 to 1 from 2 s 0 to 4 from 5 s 0 to 4 from 5 s
Release on ramped-curve from: Release on ramped-curve from: Release on ramped-curve from:
Discontinuity release detail 1 to 2 s 4 to 5 s 4 to 5 s
Runtime of simulation on 2.8 GHz processor with
9 90 145
512 Mb RAM (min)
Mesh density at boundaries 10 10 10
Mesh density applied to
5 5 5
discontinuities
Mesh spheres 5 5 5
Meshing Algorithm Delaunay Delaunay Delaunay
Geometry and
mesh detail
150 m

150 m

150 m

Mesh screen-shot illustrating


model extent

Type of model Failure due to gravity Failure due to gravity Failure due to gravity

More kinetically stable, however Finer mesh and slight change in


blocky-type fracture of failure parameters providing a more realistic
Kinetically unstable (large-scale
Result blocks prevents simulation of behaviour. Fracturing occurs
fracture following failure)
mechanism. Also fracture starts following discontinuity release
from 2.4 s (during gravity load) (at 5.1 s)

Additional comments
309
SECTION OF DATABASE: iv
Particular model (referred to in thesis): Delabole model 7 Delabole model 8 Delabole model 9
Simulation of failure mechanism Simulation of failure mechanism Simulation of failure mechanism
Model Intent
along step-path along step-path along step-path

ELFEN Analysis Version 3.8.5 3.8.5 3.8.5


Additional 'option(s)' - see Appendix A 37 166 and 37 166 and 37
K-ratio Single surface model 0.33 0.33
Surface Domain 
Single surface model Single surface model Single surface model

Control  Parameter 
(embedded discontinuities) (embedded discontinuities) (embedded discontinuities)
Value Value Value
Mass-strength and related properties

Young’s Modulus (GPa ) 20 20 20


Main
Poisson's ratio 0.25 0.25 0.25
properties
Density (kg/m3) 2500 2500 2500
Cohesion (MPa)
Mohr-
Friction Angle (°) Purely Rankine Purely Rankine Purely Rankine
Coulomb and
Dilation angle (°)
Rotating
Tensile strength (MPa) 2 2 2
crack
Fracture energy (J/m2) 20 20 20
Porosity (%)
Density of fluid (kg/m 3)
3
Groundwater Density of solid (kg/m ) Dry model Dry model Dry model
Bulk modulus of fluid (GPa)
Bulk modulus of solid (GPa)
Viscosity (Pa s)
Contact damping (%) 10 10 10
Default discontinuity and mesh-

Field (m) 0.02 0.01 0.2


Normal penalty (GPa/m) 0.1 0.1 0.1
Discrete Tangential penalty (GPa/m) 0.1 0.1 0.01
based properties

global Zone (m) 2 1 8


properties Smallest element (m) 10 10 15
Friction (°) 31 31 31
Cohesion (MPa) 0 0 0
Particular discontinuity  - - -
Particular Friction Angle (°) - - -
discontinuity Cohesion (MPa) - - -
strength Normal penalty (GPa/m) - - -
- - -
Number of stages 
Tangential penalty (GPa/m)

Stage Number  Stage Number  Stage Number 


1 2 2

1 1 2 1 2
Model Sequence

Point damping (%) 20 20 20 20 20


Critical Time step 0.75 0.75 0.25 0.75 0.25
Global gravity Global gravity
Process Global gravity loading Failure Failure
loading loading
Load/release curve Ramp-curve Ramp-curve - Ramp-curve -
Duration of stage (s) 0 to 4 0 to 4 from 5 s 0 to 4 from 5 s
Release on ramped-curve from: Release on ramped-curve from: Release on ramped-curve from:
Discontinuity release detail
4 to 5 s 4 to 5 s 4 to 5 s
Runtime of simulation on 2.8 GHz processor with
18 380 120
512 Mb RAM (min)
Mesh density at boundaries 30 30 30
Mesh density applied to
1 and 5* 1, 5 and 10* 1 and 5*
discontinuities
Mesh spheres 0 0 0
Meshing Algorithm Delaunay Delaunay Delaunay
Geometry and
mesh detail
400m

400m

400m

Mesh screen-shot illustrating


model extent

Type of model Failure due to gravity Failure due to gravity Failure due to gravity

Simulation of failure mechanism Model demonstrated chisel effect, Development of model to create a
Result although limited separation of however penetration prevents model which provided a realistic
chiselling block rotation failure mechanism

* A mesh density of 1 m was


applied to discontinuities defining * As in model 7; but also a 10 m
Additional comments failure blocks, and 5 m mesh mesh was applied to the shortah * As in model 8
density was applied to extended furthest from the face
portion of claylodes
310
SECTION OF DATABASE: iv
Particular model (referred to in thesis): Delabole model 10 Delabole model 11 Delabole model 12 Delabole model 13
Simulation of failure Plane failure model to test Step-path model to test Extended fracture model
Model Intent mechanism along step- model response to model response to to test model response to
path groundwater groundwater groundwater
ELFEN Analysis Version 3.8.5 3.9.1 3.9.1 3.9.1
Additional 'option(s)' - see Appendix A 166 and 37 166 and 37 166 and 37 166 and 37
K-ratio 0.33 0.33 0.33 0.33
Surface Domain 
Single surface model Single surface model
As in model 9 As in model 12
Control  Parameter 
(embedded discontinuity) (embedded discontinuity)
Value Value Value Value
Mass-strength and related properties

Young’s Modulus (GPa ) 20 20 20 20


Main
Poisson's ratio 0.25 0.25 0.25 0.25
properties
Density (kg/m3) 2500 2500 2500 2500
Cohesion (MPa) 2 2 2
Mohr-
Friction Angle (°) Purely Rankine 50 45 45
Coulomb and
Dilation angle (°) 5 5 5
Rotating
Tensile strength (MPa) 2 5 1.5 1.5
crack
Fracture energy (J/m2) 20 20 20 20
Porosity (%) 1 1 1
Density of fluid (kg/m 3) 1000 1000 1000
3
Groundwater Density of solid (kg/m ) Dry model 2485 2485 2485
Bulk modulus of fluid (GPa) 0.2 0.2 0.2
Bulk modulus of solid (GPa) 31 31 31
Viscosity (Pa s) 0.001 0.001 0.001
Contact damping (%) 10 10 10 10
Default discontinuity and mesh-

Field (m) 0.2 0.4 0.4 0.4


Normal penalty (GPa/m) 0.1 0.02 30* 30*
Discrete Tangential penalty (GPa/m) 0.01 0.002 10* 10*
based properties

global Zone (m) 8 2 2 2


properties Smallest element (m) 15 2 5 5
Friction (°) 31 Varied between 43 and 60° Varied between 36 and 38°
Cohesion (MPa) 0 0 1 1
Particular discontinuity - - - claylodes
Particular Friction Angle (°) - - - 45
discontinuity Cohesion (MPa) - - - 1
strength Normal penalty (GPa/m) - - - 30
- - - 10
Number of stages 
Tangential penalty (GPa/m)

Stage Number  Stage Number  Stage Number  Stage Number 


2 3* 3** 3**

1 2 1 2 1*** 2 1*** 2
Model Sequence

Point damping (%) 20 20 5 5 5 5 5 5


Critical Time step 0.75 0.25 0.75 0.25 0.75 0.25 0.75 0.25
Global Release of Release of Release of
Global gravity Global gravity Global gravity
Process gravity Failure slope slope slope
loading loading loading
loading constraint constraint constraint
Load/release curve Ramp-curve - Drop load Sine-curve Drop load Sine-curve Drop load Sine-curve
Duration of stage (s) 0 to 4 from 5 s - 1 to 5 s - 1 to 5 s - 1 to 5 s
Release on ramped- Release on ramped-curve Release on ramped-curve Release on ramped-curve
Discontinuity release detail from: 0 to 0.5 s
curve from: 0 to 0.5 s from: 0 to 0.5 s from: 0 to 0.5 s
Runtime of simulation on 2.8 GHz processor with
130 6 15 40
512 Mb RAM (min)
Mesh density at boundaries 30 30 50 50
Mesh density applied to 5 m to those defining the
1 and 5* 5 5
discontinuities failure blocks; 10 m to
Mesh spheres 3 0 2 2
Meshing Algorithm Delaunay Advancing Front Advancing Front Advancing Front
Geometry and
mesh detail
Mesh screen-shot
illustrating model extent
(all 400 m in height)

Type of model Failure due to gravity Failure due to groundwater


Successful failure due to
Introduction of finer
rise in pore pressure,
mesh (within mesh
although realistic failure
spheres) and rock-bridge Successful failure due to rise Successful failure due to
Result block interaction was not
within the shortah, allows in pore pressure rise in pore pressure
achieved due to
the simulation of failure
penetration and limited
through the mass
release

* After failure to enable unhindered sliding Pn and Pt


* Stage 3 includes the rise in
were reduced to 0.02 and 0.002 GPa/m and the
groundwater following a
Additional comments * As in Model 8 frictional strength of the base plane and shortah
ramped-curve from
reduced to 36°. ** As in Model 11. *** During this stage
6 to 7 s.
only, 10% Rayleigh damping was used.
311
LIST OF REFERENCES

List of References

ALEJANO, L. R. (2004) Failure mechanisms and design considerations on footwall slopes


with UDEC. 1st International UDEC/3DEC Symposium: Numerical Modeling of
Discrete Materials in Geotechnical Engineering, Civil Engineering and Earth
Science. ITASCA. Conference Presentations CD-Rom. Session 8. Presentation 31.
ALZO'UBI, A., MARTIN, C. D. & CRUDEN, D. M. (2007) A discrete element damage
model for rock slopes. IN EBERHARDT, E., STEAD, D. & MORRISON, T. (Eds.)
1st Canada-U.S. Rock Mechanics Symposium: Meeting Society's Challenges and
Demands. Vancouver, Taylor & Francis/Balkema. 1: Fundimentals, New
Technologies & New Ideas.
ATKINSON B.K. (1987). Fracture mechanics of rock. London: Academic Press. pp.534.
BARLA, G., BARLA, M., CHIAPPONE, A., RABAGLIATI, U. & REPETTO, L. (2003)
Continuum and discontinuum modelling of a high rock cut. 10th Congress of the
ISRM - Technology Roadmap for Rock Mechanics. South Africa, The South African
Institute of Mining and Metallurgy. pp.79-84.
BARTON N. (1972). A model study of rock-joint deformation. Int. J. Rock Mech. Min. Sci.
Vol. 9. pp. 579-582.
BEER, A. J. & MORRONGIELLO, M. (2007) Wall Optimisation - A Case Study From the
Fimiston Open Pit, Kalgoorlie Consolidated Gold Mines1. IN POTVIN, Y. (Ed.)
Slope Stability 2007 - Proceedings of the 2007 International Symposium on Rock
Slope Stability in Open Pit Mining and Civil Engineering. Perth, ACG - Australian
Centre for Geomechanics, Nedlands, Western Australia.
BOARD, M., CHACÓN, E., VARONA, P. & LORIG, L. (1996) Comparative analysis of
toppling behaviour at Chuquicamata open-pit mine, Chile. Transactions of the
Institution of Mining and Metallurgy, 105, pp.11-21.
BRACE, W. F. (1980) Permeability of crystalline and argillaceous rocks. International
Journal of Rock Mechanics & Mining Sciences Abstracts, 17, pp.241-251.
BRUMMER, R. K., LI, H. & MOSS, A. (2006) The transition from open pit to underground
mining: An unusual slope failure mechanism to Palabora. International Symposium
on Stability of Rock Slopes in Open Pit Mining and Civil Engineering. Cape Town,
The South African Institute of Mining and Metallurgy. S44, pp.411-420.
CAI, M. & KAISER, P. K. (2004) Numerical simulation of the Brazilian test and the tensile
strength of anisotropic rocks and rocks with pre-existing cracks. Int. J. Rock Mech.
Min. Sci., 41, pp.6.
CAI, M. & KAISER, P. K. (2007) Obtaining modeling parameters for engineering design by
rock mass characterization. IN RIBEIRO E SOUSA, L., OLALLA, C. &
GROSSMANN, N. F. (Eds.) 11th Congress of the International Society for Rock
Mechanics - The Second Half Century of Rock Mechanics. Lisbon, Portugal, Taylor
& Francis/Balkema. 1, pp.381-384.
CHUNG, S.-K. (2007) General Report - Theme 3: Slopes foundations and open pit mining.
IN RIBEIRO E SOUSA, L., OLALLA, C. & GROSSMANN, N. F. (Eds.) 11th
Congress of the International Society for Rock Mechanics - The Second Half
Century of Rock Mechanics. Lisbon, Portugal, Taylor & Francis/Balkema. 3,
pp.1463-1468.
COGGAN, J. S. & PINE, R. J. (1996) Application of distinct-element modelling to assess
slope stability at Delabole slate quarry, Cornwall, England. Trans. Instn. Min.
Metall., 105, pp.A22-30.

312
LIST OF REFERENCES

COGGAN, J. S., PINE, R. J., STEAD, D. & RANCE, J. R. (2003) Numerical modelling of
brittle rock failure using a combined finite-discrete element approach: implications
for rock engineering design. ISRM 2003 Technology roadmap for rock mechanics
South Africa, Published by The South African Institute of Mining and Metallurgy.
S33, pp.211-218.
COGGAN, J. S., PINE, R. J., STYLES, T. D. & STEAD, D. (2007) Application of Hybrid
Finite/Discrete Element Modelling for Back-Analysis of Rock Slope Failure
Mechanisms. IN POTVIN, Y. (Ed.) Slope Stability 2007 - Proceedings of the 2007
International Symposium on Rock Slope Stability in Open Pit Mining and Civil
Engineering. Perth, ACG - Australian Centre for Geomechanics, Nedlands,
Western Australia. pp.267-277.
CSIRO (2008a) Mine Environment Imaging Group CSIRO Division of Exploration &
Mining. Queensland, Australia 4069. http://www.em.csiro.au
CSIRO (2008b) Large Open Pit - Slope Design research [online]. [Accessed 20th July
2008]. Available from World Wide Web:
http://www.lop.csiro.au/text/Design%20Research.html
CUNDALL, P. A. (2008) Recent advances in numerical modelling for large-scale mining
projects. Australian Centre for Geomechanics Newsletter. Available from World
Wide Web:
http://www.acg.uwa.edu.au/__data/page/2139/ACGnews_June_08_lores.pdf June
ed. 30, pp.1-7.
CURRAN, J. H. & OFOEGBU, G. I. (1993) Modelling discontinuities in Numerical Analysis.
IN HUDSON, J. A. (Ed.) Comprehensive rock engineering. Pergamon. pp.443-468
DIEDERICHS, M. S., LATO, M., HAMMAH, R. E. & QUINN, P. (2007) Shear Strength
Reduction (SSR) approach for slope stability analysis. IN EBERHARDT, E.,
STEAD, D. & MORRISON, T. (Eds.) 1st Canada-U.S. Rock Mechanics
Symposium: Meeting Society's Challenges and Demands. Vancouver, Taylor &
Francis/Balkema. 1: Fundimentals, New Technologies & New Ideas, pp.319-327.
DIGHT, P. M. (2006) Pit wall failures on ‘unknown’ structures. The South African Institute
of Mining and Metallurgy. International Symposium on Stability of Rock Slopes in
Open Pit Mining and Civil Engineering. S44, pp.11-30.
DUPERRET A., GENTER A., MARTINEZ A. & MORTIMORE R.N. (2004). Coastal chalk
cliff instability in NW France: role of lithology, fracture pattern and rainfall. Coastal
Chalk Cliff Instability. Mortimore R.N. & Duperret A. (Eds.). Geological Society,
Engineering Geology Special Publication, 20, 75-88.
DURAN, A. & DOUGLAS, K. (2000) Experience with Empirical Rock Slope Design.
13/05/08. Available from World Wide Web:
http://lib.hpu.edu.cn/comp_meeting/ICGGE%E5%9B%BD%E9%99%85%E5%9C%
B0%E8%B4%A8%E5%B7%A5%E7%A8%8B%E4%BC%9A%E8%AE%AE%E8%
AE%BA%E6%96%87%E6%95%B0%E6%8D%AE%E5%BA%93/PAPERS/SNES/
SNES1186.PDF. GeoEng2000 - An International Conference on Geotechnical &
Geological Engineering. Melbourne, Australia.
EBERHARDT, E. (2003) Rock Slope Stability Analysis – Utilization of Advanced Numerical
Techniques. Earth and Ocean Sciences at UBC. Course Notes.
http://www.eos.ubc.ca/personal/erik/e-papers/EE-SlopeStabilityAnalysis.pdf.
EBERHARDT, E., STEAD, D. & COGGAN, J. S. (2004) Numerical analysis of initiation
and progressive failure in natural rock slopes - the 1991 Randa rockslide. Int. J.
Rock Mech. Min. Sci., 41, pp.69-87.
EDELBRO, C. (2003) Rock Mass Strength – A Review. Luleä University of Technology,
Sweden. 29/11/06. Available from World Wide Web: http://epubl.luth.se/1402-
1536/2003/16/LTU-TR-0316-SE.pdf

313
LIST OF REFERENCES

ELMO, D. (2006) Evaluation of a hybrid FEM/DEM approach for determination of rock


mass strength using a combination of discontinuity mapping and fracture
mechanics modelling, with particular emphasis on modelling of jointed pillars. Ph.D.
Thesis. Camborne School of Mines, University of Exeter.
ELMO, D., COGGAN, J. S. & PINE, R. J. (2005) Characterisation of rock mass strength
using a combination of discontinuity mapping and fracture mechanics modelling.
The 40th U.S. Symposium on Rock Mechanics (USRMS): Rock Mechanics for
Energy, Mineral and Infrastructure Development in the Northern Regions.
Anchorage, Alaska, American Rock Mechanics Association.
ELMO, D., VYAZMENSKY, A., STEAD, D. & RANCE, J. R. (2007a) A hybrid FEM/DEM
approach to model the interaction between open-pit and underground block-cave
mining. IN EBERHARDT, E., STEAD, D. & MORRISON, T. (Eds.) 1st Canada-U.S.
Rock Mechanics Symposium: Meeting Society's Challenges and Demands.
Vancouver, Taylor & Francis/Balkema. pp.1287-1294.
ELMO, D., YAN, M., STEAD, D. & ROGERS, S. F. (2007b) The Importance of Intact Rock
Bridges in the Stability of High Rock Slopes - Towards a Quantification
Investigation Using an Integrated Numerical Modelling; Discrete Fracture Network
Approach. IN POTVIN, Y. (Ed.) Slope Stability 2007 - Proceedings of the 2007
International Symposium on Rock Slope Stability in Open Pit Mining and Civil
Engineering. Perth, ACG - Australian Centre for Geomechanics, Nedlands,
Western Australia. pp.253-266.
ELMO, D., VYAZMENSKY, A., STEAD, D. & RANCE, J. R. (2008) Numerical analysis of
pit wall deformation induced by block-caving mining: A combined FEM/DEM - DFN
synthetic rock mass approach. IN SCHUNNESSON, H. & NORDLUND, E. (Eds.)
5th International Conference and Exhibition on Mass Mining. Sweden, Luleå
University of Technology Press. pp.1073-1082.
FAKHIMI, A. (2004) Application of slightly overlapped circular particles assembly in
numerical simulation of rocks with high friction angles. Engineering Geology, 74,
pp.129-138.
FENG, X. T. & HUDSON, J. A. (2004) The ways ahead for rock engineering design
methodologies. International Journal of Rock Mechanics & Mining Sciences, 41,
pp.255–273.
FLYNN, Z. N. (2001) Dynamic Characterisation of the caving process around longwall coal
mines using integrated microseismology and numerical modelling. University of
Liverpool. pp.138.
FLYNN, Z. N. & PINE, R. J. (2007) Fracture characterisation determined by numerical
modelling analysis. IN RIBEIRO E SOUSA, L., OLALLA, C. & GROSSMANN, N. F.
(Eds.) 11th Congress of the International Society for Rock Mechanics - The
Second Half Century of Rock Mechanics. Lisbon, Portugal, Taylor &
Francis/Balkema.
FLYNN, Z. N., PINE, R. J., COGGAN, J. S., GWYNN, X. P. & STYLES, T. D. (2007) A
new Approach for rock mass characterisation and stability analysis: A case study
using the Brighton Chalk Cliffs, UK. Poster of current research at Camborne School
of Mines, University of Exeter. Presented at Geological Society, London.
FORD, N. T. (2008) Discrete Fracture Network Modelling for the use in Block Cave Design
and Assessment. Ph.D. Thesis. Camborne School of Mines, University of Exeter.
FRANZ, J., CAI, Y. & HEBBLEWHITE, B. (2007) Numerical modelling of composite large
scale rock slope failure mechanisms dominated by major geological structures. IN
RIBEIRO E SOUSA, L., OLALLA, C. & GROSSMANN, N. F. (Eds.) 11th Congress
of the International Society for Rock Mechanics - The Second Half Century of Rock
Mechanics. Lisbon, Portugal, Taylor & Francis/Balkema. 1, pp.633-636.

314
LIST OF REFERENCES

GEOSTRU (2006) GeoStru Software [online]. [Accessed 28th November 2006]. Previously
available from World Wide Web:
http://www.geostru.com/Help/Slope/en/Cenni_teorici.htm
GOLDER (2008) FracMan Technology Group, Golder Associates (UK).
http://fracman.golder.com
GRENON, M., HADJIGEORGIOU, J. & COTE, P. (2007) Slope Stability Consideration in
Integrated Surface Mine Design. IN POTVIN, Y. (Ed.) Slope Stability 2007 -
Proceedings of the 2007 International Symposium on Rock Slope Stability in Open
Pit Mining and Civil Engineering. Perth, ACG - Australian Centre for
Geomechanics, Nedlands, Western Australia. pp.77-92.
GWYNN (2008) Assessment of Remote Data Capture Systems for the characterisation of
rock fracture networks within slopes. Ph.D. Thesis. Camborne School of Mines,
University of Exeter.
HADJIGEORGIOU, J., KYRIAKOU, E. & PAPANASTASIOU, P. (2006) A road
embankment failure near Pentalia in southwest Cyprus. International Symposium
on Stability of Rock Slopes in Open Pit Mining and Civil Engineering. Cape Town,
The South African Institute of Mining and Metallurgy. S44, pp.343-352.
HAMMAH, R. E., YACOUB, T., CORKUM, B., WIBOWO, F. & CURRAN, J. H. (2007)
Analysis of blocky rock slopes with finite element Shear Strength Reduction
analysis. IN EBERHARDT, E., STEAD, D. & MORRISON, T. (Eds.) 1st Canada-
U.S. Rock Mechanics Symposium: Meeting Society's Challenges and Demands.
Taylor & Francis/Balkema. 1: Fundimentals, New Technologies & New Ideas,
pp.319-327.
HAMMAN, E. C. F. & COULTHARD, M. A. (2007) Developing a Numerical Model for a
Deep Open Pit. IN POTVIN, Y. (Ed.) Slope Stability 2007 - Proceedings of the
2007 International Symposium on Rock Slope Stability in Open Pit Mining and Civil
Engineering. Perth, ACG - Australian Centre for Geomechanics, Nedlands,
Western Australia. pp.225-237.
HANDLEY, M. F. & KARPAROV, K. N. (2007) Proposed thrust failure analytical method for
slope collapse in open pit mines. IN RIBEIRO E SOUSA, L., OLALLA, C. &
GROSSMANN, N. F. (Eds.) 11th Congress of the International Society for Rock
Mechanics - The Second Half Century of Rock Mechanics. Lisbon, Portugal, Taylor
& Francis/Balkema. 1, pp.645-652.
HARRIES, N., NOON, D. & ROWLEY, K. (2006) Case Studies of slope stability radar used
in open cut mines. The South African Institute of Mining and Metallurgy.
International Symposium on Stability of Rock Slopes in Open Pit Mining and Civil
Engineering. S44, pp.335-342.
HARRISON, J. P. & HUDSON, J. A. (2003) Matching numerical methods to rock
engineering problems: a summary of key issues. 10th Congress of the ISRM
Technology Roadmap for Rock Mechanics. South Africa. Lisbon, Portugal, South
African Institute of Mining and Metallurgy.
HART, R. D. (1991) Introduction to Distinct Element Modelling for Rock Engineering. IN
HUDSON, J. A. (Ed.) Comprehensive Rock Engineering. Pergamon. pp.245-261
HCITASCA (2008) HCItasca Consulting Group inc. http://www.itascacg.com
HENCHER, S. & KNIPE, R. (2007) Development of joints with time and consequences for
engineering. IN RIBEIRO E SOUSA, L., OLALLA, C. & GROSSMANN, N. F. (Eds.)
11th Congress of the International Society for Rock Mechanics - The Second Half
Century of Rock Mechanics. Lisbon, Portugal, Taylor & Francis/Balkema. 1,
pp.223-226.
HOEK, E. (2004) Numerical Modelling for Shallow Tunnels in Weak Rock. Discussion
Paper 3. Rocscience. 05/12/06. Available from World Wide Web:
http://www.rocscience.com/hoek/pdf/numerical%20modelling%20of%20shallow%2
0tunnels.pdf

315
LIST OF REFERENCES

HOEK, E. (2007). Practical Rock Engineering. 20/12/08. Available from World Wide Web:
http://www.rocscience.com/hoek/pdf/Practical_Rock_Engineering.pdf
HOEK, E. & MARINOS, P. (2006) A brief history of the development of the Hoek-Brown
failure criterion. 01/12/08. Available from World Wide Web:
http://www.rocscience.com/downloads/rocdata/WebHelp/pdf_files/theory/Hoek-
Brown_History.pdf
HOEK, E. & BRAY, J. W. (1981) Rock Slope Engineering. 3rd ed. London: The Institute of
Mining and Metallurgy.
HOEK, E., KAISER, P. K. & BAWDEN, W. F. (1998) Support of Underground Excavations
in Hard Rock. Rottedam: A.A.Balkema.
HOEK, E. T., GRABINSKY, M. W. & DIEDERICHS, M. S. (1990) Numerical modelling for
underground excavation design. Trans. Instn. Min. Metall., 100, pp.A22-A30.
HOEK, E., CARRANZA-TORRES, C. & CORKUM, B. (2002) Hoek-Brown Failure Criterion
– 2002 Edition. 13/05/08. Available from World Wide Web:
http://www.rocscience.com/library/pdf/RL_1.pdf
HUANG, T.H., CHANG, C.S. & CHAO, C.Y. (2002). Experimental and mathematical
modeling for fracture of rock joint with regular asperities. Engineering Fracture
Mechanics. 69, pp.1977-1996.
ITASCA (2008) Itasca Consulting Group inc. www.itasca.ca
JING, L. (2003) A review of techniques, advances and outstanding issues in numerical
modelling for rock mechanics and rock engineering. Int. J. Rock Mech. Min. Sci.,
40, pp.283-353.
JING, L. & HUDSON, J. A. (2002) Numerical methods in rock mechanics. Int. J. Rock
Mech. Min. Sci., 39, pp.409-427.
JOHNSON, J. D., FERGUSSON, D. & GUY, G. (2007) Risk based Slope Design for
Operational Coal Mines at Rotowaro, Huntly, New Zealand. IN POTVIN, Y. (Ed.)
Slope Stability 2007 - Proceedings of the 2007 International Symposium on Rock
Slope Stability in Open Pit Mining and Civil Engineering. Perth, ACG - Australian
Centre for Geomechanics, Nedlands, Western Australia. pp.157-170.
KARAMI, A., GREER, S. & BEDDOES, R. (2008) Numerical Assessment of Step-Path
Failure of Northwest Wall of A154 Pit, Diavik Diamond Mines. IN POTVIN, Y. (Ed.)
Slope Stability 2007 - Proceedings of the 2007 International Symposium on Rock
Slope Stability in Open Pit Mining and Civil Engineering. Perth, ACG - Australian
Centre for Geomechanics, Nedlands, Western Australia. pp.293-305.
KIMBER, O. G., ALLISON, R. J. & COX, N. J. (1998) Mechanisms of failure and slope
development in rock masses. Trans Inst Br Geography, 23, pp.353–370
KLEINEI, T., LA POINTE, P. & FORSYTH, B. (1997) Realizing the Potential of Accurate
and Realistic Fracture Modeling in Mining. Int. J. Rock Mech. Min. Sci. Abstr., 34,
pp.661-661 (1).
KLERCK, P. A. (2000) The finite element modelling of discrete fracture in quasi-brittle
materials. PhD thesis. University of Swansea
KLERCK, P. A., SELLERS, E. J. & OWEN, D. R. J. (2004) Discrete fracture in quasi-brittle
materials under compressive and tensile stress states. Computer Methods in
Applied Mechanics and Engineering, 193, pp.3035-3056.
KULATILAKE, P. H. S. W., WANG, S., UCPIRTI, H. & STEPHANSSON, O. (1995) Effects
on the strength and deformability of rock masses. IN MYER, COOK, GOODMAN &
TSANG (Eds.) Fractured and Jointed Rock Masses. Balkema. pp.281-287.
KVELDSVIK, V., EIKEN, T., GANERØD, G. V. & RAGVIN, N. (2006) Evaluation of the
stability of the 700.000m2 Åknes rock slide based on data on movements and
ground conditions. The South African Institute of Mining and Metallurgy.
International Symposium on Stability of Rock Slopes in Open Pit Mining and Civil
Engineering. S44, pp.279-300.

316
LIST OF REFERENCES

LAWRENCE, J. Strength/general properties for chalk from knowledge and testing at


University of Brighton. [E-mail]. Message to: T D. Styles (t.d.styles@exeter.ac.uk).
26 September 2006.
LIGHTFOOT, N. & MACCELARI, M. J. (1999) Numerical Modelling of Mine workings Vol.
1 Part 1 and Part 2. CSIR Division of Mining Technology Rock Engineering
Programme.
LIU S.H., SUN D. & MATSUOKA H. (2005). On the interface friction in direct shear test.
Computeres and Geotechnics. Vol. 32. pp. 317-325
LOBAO, M. C. (2007) Differences between shear box modelling at Rockfield Software Ltd.
and Camborne School of Mines. [E-mail]. Message to: T D. Styles
(t.d.styles@exeter.ac.uk). 19 July 2006.
LORIG, L. J. (2007) Keynote Lecture - Using Numbers from Geology. IN RIBEIRO E
SOUSA, L., OLALLA, C. & GROSSMANN, N. F. (Eds.) 11th Congress of the
International Society for Rock Mechanics - The Second Half Century of Rock
Mechanics. Lisbon, Portugal, Taylor & Francis/Balkema. 3, pp.1369-1377.
MAKUSHA, G. & MINNEY, D. (2006) Managing the geotechnical risk associated with high
and low walls at Anglo Coal’s open pit mines. International Symposium on Stability
of Rock Slopes in Open Pit Mining and Civil Engineering. Cape Town, The South
African Institute of Mining and Metallurgy. S44, pp.301-320.
MAS IVARS, D., DIESMAN, N., PIERCE, M. & FAIRHURST, C. (2007) The Synthetic
Rock Mass approach - A step foward in the chartacterization of jointed rock
masses. IN RIBEIRO E SOUSA, L., OLALLA, C. & GROSSMANN, N. F. (Eds.)
11th Congress of the International Society for Rock Mechanics - The Second Half
Century of Rock Mechanics. Lisbon, Portugal, Taylor & Francis/Balkema. 1,
pp.485-490.
MERCER, K. (2007) Time dependent deformational behaviour of unsupported rock slopes.
Australian Centre for Geomechanics Newsletter. Available from World Wide Web:
http://www.acg.uwa.edu.au/__data/page/2139/ACG.news.dec073.pdf December
ed. 29. pp. 9-11.
MEYER, L. H. I. (2002) Numerical modelling of ground deformation around underground
development roadways, with particular emphasis on three-dimensional modelling of
the effects of high horizontal stress. Ph.D. Thesis. Camborne School of Mines,
University of Exeter.
MITANI, Y., ESAKI, T. & CAI, T. (2004) A numerical study about flexure toppling
phenomenon on rock slope. 1st International UDEC/3DEC Symposium: Numerical
Modeling of Discrete Materials in Geotechnical Engineering, Civil Engineering and
Earth Science. ITASCA. Conference Presentations CD-Rom. Session 7.
Presentation 26.
MOFFITT, K. M. & ROGERS, S. F. (2007) Probabilistic analysis of block stability in
underground excavations using realistic fracture network models. IN RIBEIRO E
SOUSA, L., OLALLA, C. & GROSSMANN, N. F. (Eds.) 11th Congress of the
International Society for Rock Mechanics - The Second Half Century of Rock
Mechanics. Lisbon, Portugal, Taylor & Francis/Balkema. 1, pp.391-393.
MOFFITT, K. M., ROGERS, S. F. & BEDDOES, R. J. (2007) Analysis of slope stability in
strong, fractured rock at the Diavik Diamond Mine, NWT. IN EBERHARDT, E.,
STEAD, D. & MORRISON, T. (Eds.) 1st Canada-U.S. Rock Mechanics
Symposium: Meeting Society's Challenges and Demands. Vancouver, Taylor &
Francis/Balkema. 2, pp.1245-1250.
MORTIMORE, R. N., LAWRENCE, J., POPE, D., DUPERRET, A. & GENTER, A. (2004a)
Coastal cliff geohazards in weak rock: the UK Chalk cliffs of Sussex. IN
MORTIMORE, R. N. & DUPERRET, A. (Eds.) Coastal Chalk Cliff Instability.
Geological Society, Engineering Geology Special Publication.

317
LIST OF REFERENCES

MORTIMORE, R. N., STONE, K. J., LAWRENCE, J. & DUPERRET, A. (2004b) Chalk


physical properties and cliff instability. Coastal Chalk Cliff Instability. Mortimore
R.N. & Duperret A. (Eds.). Geological Society, Engineering Geology Special
Publication, 20, 75-88.
MUNJIZA, A., BANGASH, T. & JOHN, N. W. M. (2004) The combined finite–discrete
element method for structural failure and collapse. Engineering Fracture
Mechanics, 71, pp.469–483.
OWEN, D. R. J., PIRES, F. M., DE SOUZA NETO, E. A. & FENG, Y. T. (2005)
Continuous/discrete strategies for the modelling of fracturing solids. NATO Science
Series. Sub Series III - Computer and Systems Sciences, 194, pp.230-266.
PINE, R. J. (1986) Rock Joint and Rock Mass Behaviour During Pressurised Hydraulic
Injections. PhD Thesis. Camborne School of Mines, University of Exeter. pp.323.
PINE, R. J. (2006a) Microsoft Excel File: Jacob. Spreadsheet for single circular / non-
circular failure surfaces, Version 1.1. Camborne School of Mines, University of
Exeter.
PINE, R. J. (2006b) Microsoft Excel File: Plane_failure. Plane failure slope stability
analysis, Version 1.3. Camborne School of Mines, University of Exeter.
PINE, R. J., COGGAN, J. S., FLYNN, Z. N. & ELMO, D. (2006) The development of a new
numerical modelling approach for naturally fractured rock masses. Rock Mechanics
and Rock Engineering, 39, No. 5, 395-419.
POISEL, R., ROTH, W., PREH, A. & ANGERER, H. (2001) The Eiblschrofen rock falls
interpretation of monitoring results based on FLAC3D investigations. The Second
International FLAC Symposium - FLAC and Numerical modelling in Geomechanics.
ITASCA. Conference Presentations CD-Rom.
POTYONDY D. (2002a). Modeling solids via bonding, parts 1 and 2: A bonded-Particle
Model for Crystalline Rock. Particle Flow Code (PFC2D/3D) Training Course. First
International PFC Symposium. Conference Presentations CD-Rom.
POTYONDY D. (2002b). PFC physics and numerical implementation. Particle Flow Code
(PFC2D/3D) Training Course. First International PFC Symposium. Conference
Presentations CD-Rom.
POROPAT, G. V. & ELMOUTTIE, M. K. (2006) Structural modelling of open pit mines.
International Symposium on Stability of Rock Slopes in Open Pit Mining and Civil
Engineering. Cape Town, The South African Institute of Mining and Metallurgy.
S44, pp.125-132.
PREH, A. & ZAPLETAL, M. (2006) The perfect mesh for FLAC3D to analyse stability of
rock slopes. The Fourth International FLAC Symposium - FLAC and Numerical
modelling in Geomechanics. ITASCA. Conference Presentations CD-Rom.
Session 1.
PSM (2004) NUMERICAL ANALYSIS OF KCGM FIMISTON PIT. Report PSM774.R1.
Australia: Pells Sullivan Meynink Pty Ltd.
QUADROS, E. (2007) Hydraulic characterization of rock masses. Geomechanical
Parameter Evaluation in Rock Engineering Practice (Day Course) - 11th Congress
of the International Society for Rock Mechanics. Lisbon, Portugal.
READ, J. (2008) Detail on Siromodel. [E-mail]. Message to: T D. Styles
(tom.d.styles@gmail.com). 22 July 07 2008.
READ, J. R. L. (2007) Rock Slope Stability Research. IN POTVIN, Y. (Ed.) Slope Stability
2007 - Proceedings of the 2007 International Symposium on Rock Slope Stability in
Open Pit Mining and Civil Engineering. Perth, ACG - Australian Centre for
Geomechanics, Nedlands, Western Australia. pp.355-359.
READ, J. R. L. & OGDEN, A. N. (2006) Developing new approaches to rock slope stability
analyses. The South African Institute of Mining and Metallurgy. International
Symposium on Stability of Rock Slopes in Open Pit Mining and Civil Engineering.
S44, pp.3-10.

318
LIST OF REFERENCES

ROBERTSON, W. (2007) Sirovision - Use of Photogrammetry for Slope Monitoring. Slope


Monitoring Forum - 2007 International Symposium on Rock Slope Stability in Open
Pit Mining and Civil Engineering (ACG).
ROCKFIELD (2006) Information and notes from training course 12/2006. Elfen Training for
Camborne School of Mines. Swansea, Rockfield Software Ltd.
ROCKFIELD (2007) Information and notes from (groundwater application) training course
08/2007. Project Interim Report: Stability of a fractured slope subjected to a water
table rise. Prepared by Lobao, M. Swansea, Rockfield Software Ltd.
ROCKFIELD (2008) Rockfield Software Ltd. Technium, Kings Road, Prince of Wales
Dock, Swansea, SA1 8PH, UK. http://www.rockfield.co.uk/elfen.htm
ROCSCIENCE (2002) Slide: A Synopsis of Slope Stability Analysis. Slide 4.0 White Paper.
Rocscience Inc., Toronto, Ontario, Canada.
http://www.rocscience.com/library/pdf/SL_1.pdf
ROCSCIENCE (2004) A New Era in Slope Stability Analysis: Shear Strength Reduction
Finite Element Technique. RocNews.
http://www.rocscience.com/library/pdf/StrengthReduction.pdf
ROCSCIENCE (2008) Analysis and design programs for civil engineering and mining
applications. www.rocscience.com.
ROMANA, M. & VASARHELYI, B. (2007) A discussion on the decrease of unconfined
compressive strength between saturated and dry rock samples. IN RIBEIRO E
SOUSA, L., OLALLA, C. & GROSSMANN, N. F. (Eds.) 11th Congress of the
International Society for Rock Mechanics - The Second Half Century of Rock
Mechanics. Lisbon, Portugal, Taylor & Francis/Balkema. 1, pp.139-142.
ROUX, R., TERBRUGGE, P. J. & BADENHORST, F. (2006) Slope management at
Navachab Gold Mine, Namibia. The South African Institute of Mining and
Metallurgy. International Symposium on Stability of Rock Slopes in Open Pit Mining
and Civil Engineering. S44, pp.579-594.
SAINSBURY, D., POTHITOS, F., FINN, D. & SILVA, R. (2007) Three Dimensional
Discontinuum Analysis of Structurally Controlled Failure Mechanisms at the Cadia
Hill Open Pit. IN POTVIN, Y. (Ed.) Slope Stability 2007 - Proceedings of the 2007
International Symposium on Rock Slope Stability in Open Pit Mining and Civil
Engineering. Perth, ACG - Australian Centre for Geomechanics, Nedlands,
Western Australia. pp.307-320.
SCHELDT, T., LU, M. & MYRVANG, A. (2003) Comparison of continuous and
discontinuous modelling for computational rock mechanics. 10th Congress of the
ISRM Technology Roadmap for Rock Mechanics. South Africa, The South African
Institute of Mining and Metallurgy.
SCHELLMAN, M. G., SEPULVEDA, R. P. & KARZULOVIC, A. (2006) Slope design for the
east wall of Mantovrede Mine, Chañaral, Chile. The South African Institute of
Mining and Metallurgy. International Symposium on Stability of Rock Slopes in
Open Pit Mining and Civil Engineering, S44, 451-470.
SIDDALL, R. G. & GALE, W. J. (1992) Strata Control – A New science for an Old Problem.
Trans. Instn. Min. Metall., 101, pp.A1-A12.
SIMMONS, J. V. & SIMPSON, P. J. (2007) Extension, Stress, and Composite Failure in
Bedded Rock Masses. IN POTVIN, Y. (Ed.) Slope Stability 2007 - Proceedings of
the 2007 International Symposium on Rock Slope Stability in Open Pit Mining and
Civil Engineering. Perth, ACG - Australian Centre for Geomechanics, Nedlands,
Western Australia. pp.213-223.
SJÖBERG, J. (1999) Analysis of Failure Mechanisms in High Rock Slopes. Doctoral
Thesis. Department of Civil Engineering, Division of Rock Mechanics. Luleå
University of Technology.

319
LIST OF REFERENCES

STACEY, P. (2006) Factors in the Design of Open Pit Slopes - A Reviewer’s Perspective.
International Symposium on Stability of Rock Slopes in Open Pit Mining and Civil
Engineering. Cape Town, The South African Institute of Mining and Metallurgy.
S44, pp.1-2.
STACEY, T. R. (2007) Slope Stability in High Stress and Hard Rock Conditions. IN
POTVIN, Y. (Ed.) Slope Stability 2007 - Proceedings of the 2007 International
Symposium on Rock Slope Stability in Open Pit Mining and Civil Engineering.
Perth, ACG - Australian Centre for Geomechanics, Nedlands, Western Australia.
pp.187-200.
STACEY, T. R., XIANBIN, Y., ARMSTRONG, R. & KEYTER, G. J. (2003) New slope
stability considerations for deep open pit mines. The Journal of The South African
Institute of Mining and Metallurgy, 104, pp.373-369.
STARFIELD, A. M. & CUNDALL, P. A. (1988) Towards a methodology for rock mechanics
modelling. International Journal of Rock Mechanics and Mining Science &
Geomechanics Abstracts, 25, pp.99-106.
STARFIELD, A. M. & CUNDALL, P. A. (1990) Towards a Methodology for Rock
Mechanics Modelling. Int. J. Rock Mech. Min. Sci. Abstr., 25, pp.99-106.
STEAD, D. & COGGAN, J. S. (2006) Numerical modelling of rock slopes using a total
slope failure approach. IN EVANS, S. G., SCARASCIA MUGNOZZA, G., STROM,
A. & HERMANNS, R. L. (Eds.) Landslide from Massive Rock Slope Failure.,
A.Springer, Dordrecht, Netherlands. pp.131-142.
STEAD, D., COGGAN, J. S. & EBERHARDT, E. (2004) Realistic simulation of rock slope
failure mechanisms: the need to incorporate principles of fracture mechanics.
SINOROCK2004, International Symposium on Rock Mechanics: Rock
characterization, modelling and engineering design methods. Three Gorges Project
site, China, Int. J. Rock Mech. Min. Sci. 41, pp.460-466.
STEAD, D., EBERHARDT, E. & COGGAN, J. S. (2006a) Developments in the
characterization of complex rock slope deformation and failure using numerical
modelling techniques. Engineering Geology, 83, pp.217-235.
STEAD, D., YAN, M., COGGAN, J. S. & EBERHARDT, E. (2006b) New developments in
the analysis of surface mine slopes: Implications for open pit slope design. 59th
Canadian Geotechnical Conference and 7th Joint CGS-IAH-CNC Groundwater
Speciality Conference, “Sea to Sky 2006”. Vancouver.
STEAD, D., COGGAN, J. S., ELMO, D. & YAN, M. (2007) Modelling Brittle Fracture in
Rock Slopes - Experience Gained and Lessons Learned. IN POTVIN, Y. (Ed.)
Slope Stability 2007 - Proceedings of the 2007 International Symposium on Rock
Slope Stability in Open Pit Mining and Civil Engineering. Lisbon, ACG - Australian
Centre for Geomechanics, Nedlands, Western Australia. pp.239-252.
SULLIVAN, T. D. (2007) Keynote Address: Hydromechanical Coupling and Pit Slope
Movements. IN POTVIN, Y. (Ed.) Slope Stability 2007 - Proceedings of the 2007
International Symposium on Rock Slope Stability in Open Pit Mining and Civil
Engineering. Perth, ACG - Australian Centre for Geomechanics, Nedlands,
Western Australia. pp.3-43.
TANG, C. A., YANG, W. T., FU, Y. F. & XU, X. H. (1998) A new approach to numerical
method of modelling geological processes and rock engineering problems --
continuum to discontinuum and linearity to nonlinearity. Engineering Geology, 49,
pp.207-214.
VAN DER PLUIJM, B.A. & MARSHAK, S. (1997) Earth Structure – An Introduction to
Structural Geology and Tectonics. USA: WCB/McGraw-Hill.
WALTHAM, A. C. (1996) 24: Rock Strength and 25: Rock Mass Strength. In: Foundations
of Engineering Geology. Glasgow, Blackie Academic & Professional. pp.48-51

320
LIST OF REFERENCES

WINES D.R. & LILLY P.A. (2003) Estimates of rock joint shear strength in part of the
Fimiston open pit operation in Western Australia. Int. J. Rock Mech. Min. Sci., 40, pp.
929-937.
WHITTAKER, B. N., SINGH, R. N. & SUN, G. (1992) Rock fracture mechanics: principles,
design and applications. Amsterdam: Elsevier. pp. 570.
WILLIAMS R.B.G., ROBINSON D.A., DORNBUSCH A., FOOTE Y.L.M., MOSES C.A. &
SADDLETON P.R. (2004) A Sturzstrom-like cliff fall on the Chalk coast of Sussex.
Coastal Chalk Cliff Instability. Mortimore R.N. & Duperret A. (eds.). Geological
Society, Engineering Geology Special Publication, 20, pp.88-97.
WYLLIE, D. C. & MAH, C. W. (2004) Rock Slope Engineering: civil and Mining. 4th ed.
London: Spon Press, Taylor & Francis Group. pp. 431.
YAN, M., ELMO, D. & STEAD, D. (2007a) Characterization of step-path failure
mechanisms: A combined field based-numerical modeling study. IN EBERHARDT,
E., STEAD, D. & MORRISON, T. (Eds.) 1st Canada-U.S. Rock Mechanics
Symposium: Meeting Society's Challenges and Demands. Taylor &
Francis/Balkema. 1: Fundimentals, New Technologies & New Ideas, pp.493-501.
YAN, M., STEAD, D. & STURZENEGGER, M. (2007b) Step-path characterization in rock
slopes: An integrated digital imaging-numerical modeling approach. IN RIBEIRO E
SOUSA, L., OLALLA, C. & GROSSMANN, N. F. (Eds.) 11th Congress of the
International Society for Rock Mechanics - The Second Half Century of Rock
Mechanics. Lisbon, Portugal, Taylor & Francis/Balkema. 1, pp.693-696.
ZHANG Z.X. (2002). An empirical relation between mode I fracture toughness and the
tensile strength of rock. Int. J. Rock Mech. and Min. Sci. Vol. 39. pp 401-406.

321

You might also like